Abstract
The sun’s energy, though free and virtually limitless, is a largely unexploited resource, as its conversion into a storable form presents several technological challenges. A promising way of capturing and storing solar energy is in the form of “solar fuels,” in a process termed artificial photosynthesis. In a photoelectrochemical (PEC) system, the reduction of CO2 to carbon-based fuels is driven on the surface of an illuminated semiconductor electrode. Through the decades, many different classes of semiconducting materials have been studied for this purpose, to varying successes. Because of their cheap and abundant nature, semiconducting transition metal oxides are good candidates to realize this technology in an economic scale and have thus attracted considerable research attention. In this review article, the progress achieved with a specific class of metal oxides, namely, the copper ternary oxides such as copper iron oxide and copper bismuth oxide, for PEC CO2 reduction is examined. Although there have been significant advances in terms of strategies to improve the efficiency and stability of these materials, further studies are warranted to address the many challenges to PEC CO2 reduction and solar fuel production.
1 Introduction
Solar energy is a free and clean resource abundant enough to supply the world’s ever-growing needs [1]. However, due to its diffuse and intermittent nature, it needs to be stored in a transportable form to become a useful energy source [2,3]. In the previous decades, several approaches to harvesting solar power have been developed, to varying levels of technological readiness [3,4,5,6]. Among these technologies is the direct storage of the sun’s energy in the chemical bonds of a “solar fuel,” in a process resembling an artificial photosynthesis system [7,8,9,10].
Hydrogen is the simplest solar fuel that can be produced via the splitting of water [11]. However, because hydrogen is a gas at standard conditions, it needs to be compressed and/or liquefied to attain a practical volumetric energy density [12]. Liquid organics that can be produced from the reduction of carbon dioxide (CO2), such as methanol and ethanol, are easier to handle, more applicable for direct use, and more compatible with the current energy infrastructure [4,5,8,9,12,13]. From this perspective, the solar-driven reduction of CO2 to fuels as well as other useful chemicals is considered an attractive route for solar energy harvesting and for closing the anthropogenic carbon cycle [12,13,14,15,16].
Before a fully operational system capable of scalable and long-term production of solar fuels is achieved, the thermodynamic and kinetic challenges to the multi-electron reduction of CO2 need to be addressed [14,17,18]. CO2, with its two C═O bonds and linear symmetry, is an extremely stable molecule [19,20,21]. Its endergonic activation requires a significant energy input due to a structural distortion accompanying electron addition [22,23,24]. Furthermore, the many reaction pathways of the activated CO2 molecule result in a variety of products, such as formic acid (HCOOH), carbon monoxide (CO), formaldehyde (HCOH), methanol (CH3OH), and methane (CH4), making product separation difficult [14,15,16,17,18]. The reduction potentials of CO2 to these products in aqueous solution, where CO2 reduction is usually conducted, are given in Table 1. The proximity of the reduction potentials with each other and with the hydrogen evolution reaction (HER) as well as the multiple proton-coupled electron transfers with different kinetic barriers for each step make CO2 reduction a rather complicated process [25,26,27,28,29].
Reduction potentials at pH 7 of some CO2 reduction reactions in aqueous solution
Reaction | E (vs normal hydrogen electrode (NHE)) |
---|---|
|
−0.61 |
|
−0.52 |
|
−0.48 |
|
−0.38 |
|
−0.25 |
In a photoelectrochemical (PEC) system, CO2 reduction is driven on an illuminated semiconductor electrode immersed in a CO2-saturated electrolyte. The photoelectrode performs the key processes involved in solar fuel generation: absorption of solar energy, charge separation and transport, and catalysis at the surface to make and break chemical bonds [30,31,32,33,34,35]. As such, there are numerous – and sometimes conflicting – requirements for the photoelectrode: wide range of sunlight absorption, high charge mobility and carrier diffusion length, suitable band energetics, good catalytic activity and selectivity, and long-term stability [33,34,35]. Many classes of materials, such as the IV, III-V, and II-VI semiconductors, metal oxides, nitrides, oxynitrides, and chalcogenides, and carbon-based materials, have been investigated for PEC CO2 reduction, yet the desired efficiency, selectivity, and stability are far from being met, even at the laboratory scale [14,15,16,25,26,27,28,29,33,34,35,36,37,38,39,40,41]. Developing new photoelectrodes by identifying new materials and designing new architectures therefore remains to be the frontier challenge for PEC CO2 reduction.
For PEC CO2 reduction to be realized in a practical scale, the photoelectrode should be made of low-cost, earth-abundant, and non-toxic materials [3,35,42]. Copper-based oxides, generally p-type semiconductors demonstrating a broad absorption of light, are good candidates as photocathodes in this context [43,44,45,46,47,48]. Because ternary and multinary oxides provide greater flexibility in tuning their properties compared to their binary counterparts, copper ternary oxides have recently gained attention as photocathodes for CO2 reduction [49,50,51]. Although research is still lacking, there have already been promising results, making it timely to review the current progress. There have been previous reviews on copper ternary oxides, but none so far has focused on their application for PEC CO2 reduction [45,46,47,48,49,50,51].
This article provides first an overview of the PEC cell and the motivation for the investigation of copper ternary oxides as materials for PEC CO2 reduction. Then, the latest efforts to improve these materials towards high efficiency and stability photocathodes are examined. Finally, insights on the way forward for these materials, as well as for PEC CO2 reduction in general, are offered.
2 Photocathode-driven PEC cell
2.1 Overview of PEC cell
The conversion of solar energy to chemical fuel in a solid-state device requires the linking of light absorption and electrochemical functionalities [3,5,6,15,25]. One configuration is a coupled photovoltaic (PV)-electrolysis system, where the PV module performs the light absorption to produce an electric potential that will drive an electrochemical reaction in the electrolyzer. In another configuration, both functionalities are integrated in a single device based on a semiconductor and the solar energy is directly converted into the energy of the redox products. Many authors claim that the latter setup is more favorable from a fabrication and cost standpoint since there will be less stringent material requirements and fewer physical components, although there is a range of intermediate configurations where the device functionalities are partially decoupled [2,5,6,15,27,29,35].
In brief, a semiconductor is a material that possesses a small or modest region (typically between 1–3 eV) of forbidden energy levels called the band gap. When a semiconductor absorbs photons more energetic than its band gap, one of the possible interactions will be for an electron from its valence band to be excited to the conduction band, leaving a hole in the valence band. Following excitation, the photogenerated charges must effectively separate before recombination occurs. This separation can be facilitated by an electric field, such as that formed at the interface of the semiconductor and an electrolyte solution, transports minority charge carriers to the surface and the majority charge carriers to the bulk (or back contact) [52,53,54]. Accordingly, the role of the semiconductor is primarily to absorb an incident photon, generate an electron-hole pair, and facilitate its separation and transport. Interfacial charge transfer to the adsorbed species to drive the redox reaction is a separate function which may be performed by the semiconductor itself or by a co-catalyst [14,15,16,55]. A more comprehensive treatment of semiconductor physics and the electrochemical behavior of the semiconductor-electrolyte interface under illumination can be found in refs. [52–54,56,57].
The semiconductor may take the form of micrometer- to nanometer-sized particles suspended in solution (Figure 1a) or a photoelectrode, usually a thin film deposited on a transparent conducting oxide substrate in contact with the electrolyte (Figure 1b) [14,32,34]. In the wireless particulate system (commonly referred to as a “photocatalytic” cell), both reduction and oxidation reactions take place on each particle, but at different sites [14,32,58]. In contrast, the photoelectrode configuration allows for the spatial separation of the reduction and oxidation reactions and for the application of an external bias [33,34]. This system is referred to as a PEC cell (or photoelectrocatalytic/photoelectrosynthetic) as it resembles the two- or three-electrode cell construction for electrochemical systems [14,15,16,29,30,31,32,33,34,35,40,41,53]. It is important to note here that catalysis, by definition, involves a thermodynamically favorable reaction (ΔG < 0), where the catalyst only serves to lower the activation energy (E A) of the reaction. Thus, the use of the terms “photocatalytic” or “photoelectrocatalytic” in the context of CO2 reduction, a thermodynamically uphill reaction (ΔG > 0), is imprecise, as pointed out by several authors [5,25,27,59]. In this case, an energy input, which may be provided by incident photons, is used to overcome the thermodynamic barrier ΔG rather than to lower the kinetic overpotential E A. However, the semiconductor may be considered catalytic in the sense that it can also lower E A by providing active sites for faster interfacial kinetics. Thus, it is also possible for the reaction to occur at an “underpotential,” where the light energy is utilized in such a way that both ΔG and E A are compensated, resulting in a net negative ΔG of the reaction [60,61].
![Figure 1
(a) Schematic of a photocatalytic cell, where semiconductor particles are suspended in a CO2-containing electrolyte. Each particle acts as tiny photoelectrolysis cells where both reduction and oxidation reactions take place. (b) Schematic of a typical two-compartment, three-electrode cell construction of a photocathode-driven PEC cell for CO2 reduction. Absorption of light (hν) more energetic than the band gap (E
g) of the photocathode results in the generation of electron-hole pairs, separated by the electric field (as shown by the bending of the bands) at the semiconductor-electrolyte interface. Electrons are transported to the surface, where they participate in the reduction of CO2, while the holes migrate to the back contact and to the counter electrode (CE) via the external circuit, where they are consumed in the oxidation reaction (in aqueous electrolytes, the oxygen evolution reaction), completing the flow of current in the cell. The addition of a reference electrode (RE) allows the potential of the WE (or more specifically, the Fermi level of the photocathode) to be controlled via a potentiostat, which is useful in the investigation of PEC systems. (Adapted with permission from ref. [14]. Copyright © 2015, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_001.jpg)
(a) Schematic of a photocatalytic cell, where semiconductor particles are suspended in a CO2-containing electrolyte. Each particle acts as tiny photoelectrolysis cells where both reduction and oxidation reactions take place. (b) Schematic of a typical two-compartment, three-electrode cell construction of a photocathode-driven PEC cell for CO2 reduction. Absorption of light (hν) more energetic than the band gap (E g) of the photocathode results in the generation of electron-hole pairs, separated by the electric field (as shown by the bending of the bands) at the semiconductor-electrolyte interface. Electrons are transported to the surface, where they participate in the reduction of CO2, while the holes migrate to the back contact and to the counter electrode (CE) via the external circuit, where they are consumed in the oxidation reaction (in aqueous electrolytes, the oxygen evolution reaction), completing the flow of current in the cell. The addition of a reference electrode (RE) allows the potential of the WE (or more specifically, the Fermi level of the photocathode) to be controlled via a potentiostat, which is useful in the investigation of PEC systems. (Adapted with permission from ref. [14]. Copyright © 2015, American Chemical Society).
PEC cell constructions vary depending on the number of photoactive materials and reaction compartments [16,35,38,41,42]. In a single photoelectrode configuration, a photocathode (or photoanode), based on a p-type (n-type) semiconductor, drives the reduction (oxidation) half-reaction, with the other half-reaction taking place on a dark anode (cathode), usually a metal electrocatalyst (such as platinum). The basic operation of a three-electrode PEC cell based on a photocathode is illustrated in Figure 1b. A dual photoelectrode configuration is also possible, where the photocathode and photoanode may be in tandem (both semiconductors on a single substrate) or separately illuminated, provided that the semiconductors have matching band gaps and band edge positions [15,35]. Although the ultimate goal for PEC applications is an unbiased cell made of dual photoelectrodes, the photocathode and photoanode can be developed independently [35]. In fact, research on photoanode materials has progressed more and has been the subject of the majority of previous reviews [25,36,37,38,39].
2.2 Photocathode materials for CO2 reduction
PEC CO2 reduction was first reported by Halmann, who studied a p-GaP photocathode irradiated with UV light in a CO2-saturated aqueous solution, yielding formic acid (and smaller amounts of formaldehyde and methanol) [62]. Much of the early work on PEC CO2 reduction involved covalent semiconductors, such as p-GaP [63,64,65,66], p-GaAs [63,66,67,68,69], p-InP [66,67,70,71], p-CdTe [70,72,73,74], and p-Si [66,75], being modeled after the several efficient liquid-junction solar cells developed during that period. However, even if their band gaps are suitable for visible light absorption, these semiconductors required extremely negative potentials to effect CO2 reduction, implying that little to no solar energy was being stored in the products [32]. Modification techniques such as addition of metal co-catalyst [76,77,78,79], polymer coating [80,81], and use of solution mediators [82,83,84,85,86] were needed to improve their catalytic activity as well as their stability against corrosion in aqueous electrolyte. These limitations, as well as their high cost and complicated preparation methods, have motivated researchers to explore new classes of materials for PEC CO2 reduction.
For PEC technology to become economically viable, cheap and earth-abundant materials that can be synthesized using simple and low-cost methods must be used, for which semiconducting transition metal oxides are a good candidate [35,46,49,87,88]. Transition metal oxides are a class of unique materials that exhibit a variety of structures and properties. Depending on the nature of the metal d-orbital, the metal-oxygen bonding can vary between nearly ionic to highly covalent or metallic, resulting in a wide range of electronic behavior, from insulators to semiconductors to superconductors [89,90]. Semiconducting behavior of metal oxides result from intrinsic point defects acting as donor or acceptor states: oxygen vacancies for n-type oxides, while metal vacancies for p-type oxides [90]. Although the first report [91] on the use of semiconducting metal oxides for solar-driven CO2 reduction came around the same time as Halmann’s, research interest had only expanded in the succeeding years, with the advent of nanotechnology and modern computational capabilities [92]. As mentioned earlier, photoanode materials, in particular the most well-studied TiO2, have received more attention than photocathode materials [25,35,36,37,38,39,49]. This is due in part to the fact that there are fewer p-type oxides because of the localized nature of the oxygen 2p orbitals in the valence band of metal oxides that results in large effective hole masses and easier hole compensation [93]. Even so, the small band gap p-type oxides (and thus suitable as photocathodes), namely, Cu2O (∼2.0 eV) and CuO (∼1.5 eV), are susceptible to photocorrosion in aqueous electrolyte [43,44,45,46,47,48]. Strategies to enhance the stability and PEC activity of Cu2O and CuO have been demonstrated, such as nanostructuring [94,95,96], addition of co-catalyst [97,98,99,100,101,102,103], deposition of protective overlayers [104,105,106,107], and heterojunction formation [61,108,109,110,111,112,113,114,115,116], to varying successes. The performance of Cu2O- and CuO-based photocathodes for CO2 reduction has been reviewed previously [45,46,48] and will no longer be discussed in this article, but a summary of previous findings is presented in Table 2.
Cu2O- and CuO-based photocathodes for PEC CO2 reduction
Photocathode | Reaction conditions | Products formed | Photostability | Ref. |
---|---|---|---|---|
CuO–Cu2O nanorods | AM 1.5 illumination (70 mW·cm−2), CO2-saturated 0.1 M Na2SO4 | CH3OH (FE: 95%) at −0.20 V vs standard hydrogen electrode (SHE), 1.5 h | Nanorods remain intact after 2 h photoelectrolysis | [61] |
M/CuO/Cu2O (M = Ag, Au, Cd, Cu, Pb, Sn) | 450 W Xe lamp, CO2-saturated 0.1 M KHCO3 | CH3OH, CO, HCOOH (total FE: 40.45% for M = Pb) at −0.16 V vs SHE, 1 h | 70% decrease in photocurrent after 20 min due to reduction of CuO to Cu2O or Cu | [99] |
Cu/Cu2O | 125 W high pressure Hg lamp, CO2-saturated 0.1 M Na2CO3/NaHCO3 | CH3OH, CH2O, C2H5OH, CH3COH, CH3COCH3 (80% total CO2 reduced) at 0.2 V vs Ag/AgCl, 2 h | — | [100] |
Cu2O/TiO2/Re(tBu-bipy)(CO)3Cl | 450 W Xe arc lamp, CO2-saturated acetonitrile with 0.1 M TBAPF6 | CO (FE: ∼100%) at −1.73 V vs Fc/Fc+, 5.5 h | A small decrease in photocurrent was observed over 5.5 h of testing | [101] |
Cu2O/CuO coated Ag dendrites | Warm white LED (100 mW·cm−2), CO2-saturated 0.1 M Na2SO4 | CH3COO− (FE: 54%) CO (trace) at −0.4 V vs Ag/AgCl, 1 h | — | [103] |
Cu2O nanowires coated with Cu+-TiO2 | 150 W LS Xe arc lamp, CO2-saturated 0.3 M KHCO3 | CH3OH (FE: 56.5%), CO, CH4 (trace) at 0.3 V vs reversible hydrogen electrode (RHE), 2 h | 72.4% reduction in the photocurrent after 30 min of testing | [106] |
Cu2O/Cu3(BTC)2 | 300 W Xe lamp, CO2-saturated acetonitrile with 0.1 M TBAPF6 | CO (FE: ∼95% between −1.77 and −1.97 V vs Fc/Fc+, STC: 0.83% at −2.07 V vs Fc/Fc+), 1 h | Nearly constant over 3,500 s at −1.97 V vs Fc/Fc+ | [107] |
CuO clusters on Fe2O3 nanotubes | Xe lamp (100 mW·cm−2), CO2-saturated 0.1 M KHCO3 | CH3OH (FE: 91.2%), C2H5OH (9.8%), CH4 (10.46%) at −1.1 V vs SCE, 6 h | — | [108] |
Ultra-long CNT/Cu2O | Solar simulator (100 W), CO2-saturated 0.1 M Na2SO4 | CH3OH (FE: 9%), C2H5OH (24%), HCOOH (18%) at 0.05 V vs Ag/AgCl, 4 h | 40% reduction in the photocurrent after 4 h of testing | [109] |
Ti/TiO2/CuO | 125 W high pressure Hg lamp, CO2-saturated 0.1 M NaHCO3 | CH3OH (97 mol%), C2H5OH, (CH3)2CO (trace) at −0.60 V vs Ag/AgCl, 2 h | — | [110] |
CuO/g-C3N4 | Light source (51.6 W·m−2), CO2-saturated 0.1 M NaHCO3 | CH3OH (FE: 75%, QE: 8.9%) at −0.4 V vs NHE, 4 h | — | [113] |
CuO/CdS | 250 W lamp, CO2-saturated 0.1 M NaHCO3 | CH3OH (FE: 86%) at −0.4 V vs NHE | — | [114] |
CuO/Cu2O||rGO|h-WO3||rGO | 150 W solar simulator, CO2-saturated 0.1 M Na2SO4 | CH3OH (FE: ∼75%), CO (∼25%) at 0.1 V vs RHE, 1 h | rGO layer increased stability of the system, although no actual measurements were shown | [116] |
FE – Faradaic efficiency, STC – solar-to-carbon efficiency.
Another promising strategy to address the limitations of binary metal oxides is by expanding the search for photoelectrode materials to multinary metal oxides [35,42,46]. The addition of one or more cations to form more complex oxides provides greater flexibility for tuning the optical and electronic properties [42,45]. For ternary and quaternary oxides, more than 8,000 and 700,000 combinations are possible, respectively [49]. With this immense number, it is likely that a semiconductor with the best combination of properties for PEC CO2 reduction may be found among this group. Like the case with binary oxides, much of the success with ternary oxides has been achieved with photoanodes, specifically BiVO4 [35,46,49,51]. There is still a limited understanding of ternary oxide photocathodes and, even more, their application to PEC CO2 reduction.
Several copper ternary oxides exhibit p-type behavior, a narrow band gap capable of visible light absorption, and an adequate conduction band potential for CO2 reduction, as shown in Figure 2. In addition, they appear to be more stable than their binary counterparts, as the additional cation introduces modifications to the electronic structure and to the nature of band transitions that lead to inhibition of Cu self-reduction [49,50]. Along with their inexpensive and nontoxic nature, these copper ternary oxides are good candidates as photocathodes for CO2 reduction [51]. Still, several challenges limit the actual performance of copper ternary oxide photocathodes. First is the poor charge separation and transport due to low carrier mobility, a common problem among transition metal oxides [87]. Compared to covalent semiconductors, transition metal oxides are more prone to carrier localization or “self-trapping” due to a strong electron-lattice interaction which causes local lattice distortions called the small polaron. Carrier transport then occurs via a small-polaron hopping mechanism, which is characterized by a very small drift mobility due to its thermally activated nature [117,118]. Next, with the increasing complexity of metal oxides, preparation of stoichiometric, high-quality films becomes more challenging. Even a small percent of sub-stoichiometry can result in large number of defects that can act as recombination centers and diminish the photoactivity [49,51]. Lastly, although there are encouraging results, the long-term stability needed for PEC applications is yet to be demonstrated [50,51].
![Figure 2
Conduction band and valence band position of several copper ternary oxides relative to CO2 and water redox potentials, plotted versus vacuum (left) and SHE (right) scales. (Adapted from ref. [50]). Copyright © 2022, IOP Publishing.](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_002.jpg)
Conduction band and valence band position of several copper ternary oxides relative to CO2 and water redox potentials, plotted versus vacuum (left) and SHE (right) scales. (Adapted from ref. [50]). Copyright © 2022, IOP Publishing.
3 Copper ternary oxide photocathodes
In the following sections, the current progress with copper ternary oxides is summarized and critically examined. In-depth focus is given to the relatively more studied oxides, copper iron oxide (CuFeO2) and copper bismuth oxide (CuBi2O4), discussing the relation between their crystal and electronic structure, their various preparation methods, the strategies applied to address the aforementioned limitations, and their performance towards PEC CO2 reduction.
3.1 CuFeO2
CuFeO2 belongs to a group of Cu(i)-based oxides with the delafossite structure and the general formula CuMO2, where M is a metal cation in the trivalent state (M = Al, Co, Cr, Fe, Ga, and Rh) [119]. The delafossite structure comprises of two alternating layers: a planar layer of Cu cations in a triangular pattern and a layer of edge-sharing MO6 distorted octahedra, where each oxygen is coordinated by one Cu+ and three M3+ cations (Figure 3a and b). Cu and O form linear arrays parallel to the c-axis. Depending on the stacking pattern of the MO6 layers, delafossites form two polytypes, namely, 3R and 2H, that crystallize in the rhombohedral (
![Figure 3
Two polytypes of the delafossite structure: (a) 3R and (b) 2H (brown spheres – Cu, green – M, and blue – O). (Reprinted with permission from ref. [119]. Copyright © 2006, Elsevier). (c) Upon optical excitation of CuFeO2, electrons are promoted from hybridized Cu 3d and O 2p states in the valence band to Fe 3d states in the conduction band. This is followed by an ultrafast hole thermalization where electrons in the higher lying Cu 3d states backfill the photogenerated holes in the lower O 2p states. (Reprinted with permission from ref. [129]. Copyright © 2018, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_003.jpg)
Two polytypes of the delafossite structure: (a) 3R and (b) 2H (brown spheres – Cu, green – M, and blue – O). (Reprinted with permission from ref. [119]. Copyright © 2006, Elsevier). (c) Upon optical excitation of CuFeO2, electrons are promoted from hybridized Cu 3d and O 2p states in the valence band to Fe 3d states in the conduction band. This is followed by an ultrafast hole thermalization where electrons in the higher lying Cu 3d states backfill the photogenerated holes in the lower O 2p states. (Reprinted with permission from ref. [129]. Copyright © 2018, American Chemical Society).
Among the copper delafossites, CuFeO2 possesses the narrowest optical band gap (1.36‒1.55 eV), making it a suitable photocathode material for PEC applications [122,123,124]. It exhibits a p-type conductivity, which is believed to arise from native Cu vacancies and O interstitials in its structure [124,125,126]. Electronic structure calculations show that the valence band maximum in CuFeO2 is dominated by Cu 3d states, with some degree of hybridization from O 2p states, while the conduction band minimum consists of nonbonding Fe 3d states [127]. The lowest optical transition involves an electron excitation from O 2p states to Fe 3d states [128]. This excitation is followed by an ultrafast thermalization of photogenerated holes from O 2p to Cu 3d valence band states, which may be visualized as an electron transfer from Cu 3d states to backfill holes in the deeper O 2p states [129] (Figure 3c). The hole thermalization process facilitates charge separation in the lattice and suppresses electron-hole recombination (because there is no covalent bonding between the Cu and Fe atoms), and hence, is considered responsible for the greater carrier lifetime, improved stability, and superior photoactivity of CuFeO2 relative to its parent binary oxide, Cu2O [126,129]. CuFeO2 photocathodes have been prepared using a variety of methods, including hydrothermal synthesis [121,130], electrodeposition [122,126], solid-state synthesis [123,131], sol-gel method [132,133], spray-pyrolysis [134], and reactive co-sputtering [127,135].
Based on its band gap, the theoretically achievable photocurrent of CuFeO2 is 15 mA·cm−2 (based on AM1.5 illumination), but reported photocurrents are much lower (up to 2.5 mA·cm−2) [124]. Several studies have been conducted to determine the major limiting factor to the PEC performance of CuFeO2. Time-resolved microwave conductivity measurements found that charge carriers in CuFeO2 are relatively long-lived (as compared to Cu2O or α-Fe2O3), with a time constant of 200 ns [124]. On the contrary, a very short free carrier lifetime (in the order of picoseconds) was measured via time-resolved optical spectroscopy [127]. The results of these two studies indicate that free carriers generated upon illumination quickly self-trap, forming the observed long-lived excited state that is postulated to be electron thermalizing as a surface-trapped small polaron in the Fe 3d conduction band state [136]. As mentioned in the previous section, a conduction mechanism based on small polaron hopping will result in small carrier mobilities. Another factor is the presence of a high density of surface states on CuFeO2, arising from a 10 nm hydroxide or oxyhydroxide surface layer [124]. These surface states act as electron traps, causing severe Fermi level pinning and drastically reducing the photovoltage. Thus, preparation of high-quality films as well as other strategies to address the low carrier mobility and to passivate or remove the surface states is critical to improve the PEC performance of CuFeO2.
To increase its conductivity, CuFeO2 can be doped with a divalent cation (such as Mg2+, Ni2+, and Sn2+), which substitutes with either one Fe3+ or three Cu+ to contribute additional holes [123,133,137]. Indeed, the hole concentration was found to increase with Mg doping level, but only until a certain threshold where the conductivity of CuFeO2 shifts to n-type, possibly due to charge compensation by oxygen vacancies, an n-type defect [131,133,138]. A 0.05% Mg-doped CuFeO2 was studied for PEC CO2 reduction, yielding photocurrents of up to 1 mA·cm−2 in CO2-saturated 0.1 M NaHCO3 (pH 6.8) [123]. Formate was observed as the main reduction product (although no quantification was made) in bulk electrolysis experiments performed under blue LED light illumination (470 nm, 2.1 mW·cm−2). However, no comparison in the CO2 reduction activity between the doped sample and an undoped one was made. Alternatively, carrier concentration can also be increased via oxygen intercalation, which can be achieved via conventional thermal annealing [132] or hybrid microwave annealing [139] of as-synthesized CuFeO2 films. The O interstitials occupy the Cu planes, where there is a large enough space compared to the Fe octahedrons, of the delafossite structure [140]. In one study, the increase in photocurrent of the oxygen-intercalated samples versus as-synthesized CuFeO2 matched that of the acceptor density, which the authors ascribed to an improved charge separation efficiency due to the higher carrier conductivity [132]. On the contrary, another study found that O interstitials have a negative effect on PEC activity of CuFeO2 [126]. Because the O interstitials shift the Cu 3d band to lower energy and the O 2p band to higher energy, the valence band maximum changes from Cu 3d to O 2p upon introduction of the O interstitials, precluding the hole thermalization kinetics responsible for the increased carrier lifetimes in CuFeO2, as discussed above. These conflicting findings suggest the existence of a more complex underlying defect chemistry in CuFeO2 that may require further elucidation. To address the presence of surface traps in CuFeO2 and the slow interfacial kinetics, the addition of a co-catalyst or a protective overlayer has been performed. Noble metals (such as Pt, Ag, and Au) [141,142] and layered double hydroxides [139,143] were employed as co-catalyst, all resulting in an increased photocurrent. Au nanoparticles deposited on diamond-like carbon (DLC) was applied as overlayer for CuFeO2 [142]. In this strategy, the Au nanoparticles contribute to an enhanced light absorption through its surface plasmon resonance whereas the DLC support prevents agglomeration of the Au nanoparticles and decreases charge trapping surface states on CuFeO2. The authors note that further work is needed to optimize the morphology and thickness of the overlayer.
Other approaches to improve the performance of CuFeO2 include formation of novel nanostructures that enhance light absorption and charge separation. CuFeO2-coated amorphous SiO2 microspheres were assembled on an fluorine-doped tin oxide (FTO) substrate to form a monolayer opal photocathode with glass-like transparency (Figure 4a) [144]. Due to the two-dimensional photonic architecture, the photocathode demonstrated self-light harvesting. Despite the enhanced light absorption, the incident photon-to-current efficiency (IPCE) was low due to the very small thickness (∼2 nm) of the CuFeO2 shell that cannot support a depletion layer. To improve the charge separation, the CuFeO2-coated microspheres were coated with an outer shell of CuAlO2, forming a heterojunction between the two oxides [145]. It was initially shown that a host-guest architecture for CuFeO2 and CuAlO2 (a thin film of CuFeO2 on a highly transparent CuAlO2 scaffold) facilitates hole transfer, while blocking electron transfer from the absorber to the scaffold, reducing charge recombination and attaining higher photocurrents [146]. An inverse opal structure was also constructed for CuFeO2 by the same group of authors (Figure 4b) [143]. In this structure, light absorption was enhanced owing to multiple internal light scattering in the macropores. The inverse opal structure also shortened the electron path to the electrolyte, resulting in a two-fold increase in photocurrent versus that of a planar CuFeO2. In another study, Nb-doped TiO2 nanotubes were employed as substrate for CuFeO2, creating a p-n composite photocathode for the PEC reduction of CO2 [147]. Nb doping improved the heat stability of the TiO2 nanotubes, so that the structure can retain its integrity after the calcination step in the formation of CuFeO2. The authors claimed that a Z-scheme heterojunction was formed, where the weaker photogenerated electrons from TiO2 recombined with the holes from CuFeO2, allowing the more reductive CuFeO2 electrons to reduce CO2. Formaldehyde and ethanol were the CO2 reduction products detected after photoelectrolysis carried out in CO2-saturated 0.1 M NaHCO3 electrolyte under a 250 W Xe lamp illumination, with ethanol dominating at higher bias potentials.
Several reports have found that composites of CuFeO2 with binary copper oxides (Cu2O or CuO), typically resulting from nonstoichiometry of precursors during preparation, perform better than phase-pure CuFeO2 [127,148,149,150,151]. In particular, CuFeO2/CuO photoelectrodes produced much higher photocurrents than CuFeO2 or CuO alone [127,148]. These studies hypothesize that a type II heterojunction forms between the CuFeO2 and CuO phases. This heterojunction is then responsible for an increased carrier lifetime and presumably, to the observed improved photoactivity of the mixed-phase photoelectrodes. In a proposed scheme (Figure 5), photogenerated electrons from the higher E cb of CuFeO2 can migrate to CuO, whereas those from the lower E cb (an optically forbidden gap) recombine with photogenerated holes from CuO, which cannot effectively perform water oxidation as compared to holes from CuFeO2 [152]. This internal recombination allows for an improved charge separation.
![Figure 5
Electronic band diagram of CuFeO2/CuO heterojunction. (Reprinted with permission from ref. [152]. Copyright © 2019, Elsevier).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_005.jpg)
Electronic band diagram of CuFeO2/CuO heterojunction. (Reprinted with permission from ref. [152]. Copyright © 2019, Elsevier).
Based upon these findings, an unbiased CuFeO2/CuO-Pt cell under circumneutral pH was devised [148]. The cell produced formate for over a week, with a solar-to-formate efficiency of 0.7‒1.2% and Faradaic efficiency of 90%. The CuFeO2/CuO electrode, synthesized via electrodeposition, had a bulk Cu/Fe atomic ratio of 1.4 (measured via inductively coupled plasma-mass spectrometry) and a double-layered structure where CuFeO2 was located in the bottom region, while CuO was uniformly distributed throughout, as determined via transmission electron microscopy-energy-dispersive X-ray spectroscopy. Thus, CuO was segregated as a secondary phase during calcination of the precursor in the presence of air. Notably, the authors did not detect other products (or possibly, were below detection limits), especially H2 which is known to compete with CO2 reduction. For comparison, single phase CuFeO2 (Cu/Fe = 1.06) and CuO electrodes were tested under identical experimental conditions. Formate production rates were much lower, and more importantly, O2 evolution in the Pt counter electrode was observed in neither. In a follow-up study, they investigated the stability of the CuFeO2/CuO photoelectrode for 7 days of PEC testing [150]. X-ray diffraction (XRD) analysis of the used samples showed a decrease in the intensity of CuFeO2 and CuO peaks, while new peaks ascribed to Cu2O were found, indicating occurrence of cathodic photocorrosion. Reannealing of the used electrodes restores the performance similar to that of the as-synthesized samples. In another study, the effect of varying the Cu/Fe ratio on the selectivity of the CO2 photoreduction process was investigated [149]. The authors found that decreasing the Cu/Fe ratio (more Fe-rich) favors the formation of acetate, a reduction product with C‒C coupling. However, a consequence is that the Faradaic efficiency lowers (Figure 6a and b). Because minimal H2 evolution was observed in all Cu/Fe ratios, the authors attributed the Faradaic efficiency loss to self-reduction of the photocathode. X-ray photoelectron spectroscopy (XPS) of the tested samples revealed almost no surface Fe, even for the most Fe-rich sample. Correspondingly, acetate formation is observed to subside after 10 min of reaction, revealing that surface Fe atoms act as active sites for acetate formation [149].
![Figure 6
(a) Concentration of formate and acetate and (b) the corresponding Faradaic efficiencies after photoelectrolysis using CuFeO2 electrodes with various Fe:Cu atomic ratios. (Reprinted with permission from ref. [149]. Copyright © 2017, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_006.jpg)
(a) Concentration of formate and acetate and (b) the corresponding Faradaic efficiencies after photoelectrolysis using CuFeO2 electrodes with various Fe:Cu atomic ratios. (Reprinted with permission from ref. [149]. Copyright © 2017, American Chemical Society).
To provide a theoretical understanding of CO2 reduction on CuFeO2, the mechanism of CO2 adsorption and activation on CuFeO2 surface was investigated [153]. In this study, the authors found that the defect-free (011) surface was inert towards CO2 adsorption but that oxygen vacancies result in a negative charge accumulation on nearby Fe atoms, consequently serving as potential active sites for adsorption. Oxygen vacancies on many metal oxides are well known to play a role in CO2 adsorption [154]. In another study, the effect of the heterogeneity of the photoelectrode on the selectivity of the PEC CO2 reduction was investigated [152]. In their DFT calculations, the authors employed two models, a homogeneous structure (HMS) and heterogeneous structure (HTS). The former was based on a homogeneous distribution of Cu, Fe, and O in the entire film (40% CuFeO2 and 60% CuO), while the latter was based on a CuO layer (3 layers) on top of CuFeO2 (2 layers). Bader charge analysis showed that in both the HMS and HTS models, Cu atoms were more enriched with electrons than the surrounding Fe or O atoms, indicating that Cu atoms serve as active sites for the electrophilic CO2 molecule. Further analysis showed that monodentate coordination of an O atom of CO2 to a Cu site on the HTS surface was the most kinetically preferred pathway for formate production. Hence, these findings can be correlated to the higher rate of formate production with the heterogeneous CuFeO2/CuO photoelectrodes. Nevertheless, the surface Fe atoms can also act as electron rich sites; however, adsorbed CO2 on Fe sites are preferentially reduced to surface-bound *CO or *C species that blocks C‒H bond formation (a prerequisite to formate production) [155]. This can account for the acetate formation observed in Fe-rich photoelectrodes [149].
3.2 CuBi2O4
CuBi2O4 (“kusachiite”) is a p-type oxide with a tetragonal crystal structure (space group P4/ncc). It comprises square planar [CuO4]6‒ units stacked along the c-axis in a staggered manner, with Bi3+ ions positioned between the stacks and connected to six O2‒ ions with three different bond distances (Figure 7a) [156]. Accordingly, a c-axis projection of the crystal structure shows straight channels formed around Bi3+ ions (Figure 7b) [157]. The isolated [CuO4]6‒ stacks, which are unlike the edge-sharing oxygen octahedra or tetrahedra commonly found in metal oxides, define the unique crystal structure of CuBi2O4 and play a significant role in its electronic properties [158]. Initially identified as a promising photoactive material in a high-throughput screening study, CuBi2O4 possesses an optical band gap of 1.5‒1.8 eV, making it a good photocathode for PEC devices [157,159].
![Figure 7
(a) Isometric and (b) c-axis projection of the crystal structure of CuBi2O4 (Cu – orange, Bi – purple, and O – red). (Reprinted with permission from ref. [157]. Copyright © 2021, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_007.jpg)
(a) Isometric and (b) c-axis projection of the crystal structure of CuBi2O4 (Cu – orange, Bi – purple, and O – red). (Reprinted with permission from ref. [157]. Copyright © 2021, American Chemical Society).
Unlike in CuFeO2, the Cu ions in CuBi2O4 adopt a +2 oxidation state and thus, an open shell 3d9 configuration that allows intra-atomic electronic transitions [160]. Initially, the valence and conduction band edges of CuBi2O4 were considered to be primarily of O 2p and Cu 3d character, respectively. Cu vacancies, the most probable defect in CuBi2O4, introduce states near the valence band edge that gives rise to its p-type behavior [161]. However, a recent first principle and comprehensive spectroscopic characterization study revealed that the valence band maximum and conduction band minimum arise from the occupied and unoccupied spin states, respectively, of the Cu 3d x²‒y² (hybridized with O sp3 orbitals) that formed from the splitting of Cu 3d levels in the square-planar crystal field of the [CuO4]6‒ units [157]. The orbital composition of the valence and conduction bands of CuBi2O4 is depicted in Figure 8a. Owing to the strong hybridization between Cu 3d and O sp3 orbitals, Cu 3d character is present across the whole valence band up to the maximum. Contrary to initial reports that the valence band maximum has a significant Bi 6s character, Bi 6s only occurs deep into the valence band. On the other hand, beyond its minimum which consists of the unoccupied Cu 3d x²‒y² spin state, the conduction band is predominantly of Bi 6p character [157,160].
![Figure 8
(a) Orbital composition of valence and conduction bands of CuBi2O4, with the H+/H2 and O2/H2O redox potentials as reference. (Reprinted with permission from ref. [157]. Copyright © 2021, American Chemical Society). (b) Electron density difference field isosurface contours illustrating lack of overlap of electron density among [CuO4]6‒ units. (c) Electron density isosurface in b-axis projection, showing Cu‒Cu and Cu‒Bi interatomic distances and most probable direction of polaron hopping. (b and c reprinted with permission from ref. [158]. Copyright © 2020, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_008.jpg)
(a) Orbital composition of valence and conduction bands of CuBi2O4, with the H+/H2 and O2/H2O redox potentials as reference. (Reprinted with permission from ref. [157]. Copyright © 2021, American Chemical Society). (b) Electron density difference field isosurface contours illustrating lack of overlap of electron density among [CuO4]6‒ units. (c) Electron density isosurface in b-axis projection, showing Cu‒Cu and Cu‒Bi interatomic distances and most probable direction of polaron hopping. (b and c reprinted with permission from ref. [158]. Copyright © 2020, American Chemical Society).
Consequently, the lowest-energy optical absorption in CuBi2O4 corresponds to a Cu d‒d excitation, not the O 2p to Cu 3d charge transfer transition initially proposed [161]. While the Cu d‒d transition accounts for the low-energy visible light absorption of CuBi2O4, it does not significantly contribute to the photoresponse, as localized d‒d transitions are known to be less efficient in converting photons to photocurrent than O2‒ ligand to metal cation charge transfer (LMCT) transitions in open-shell metal oxides [157]. This can very well explain why even though CuBi2O4 exhibits an absorption onset at low energies (1.5‒1.8 eV), the IPCE and absorbed photon-to-current efficiency become practically significant only at higher energies (1.8‒2.25 eV), as reported in many experimental studies [156,162,163,164,165,166,167,168]. Meaningful photocurrent generation in CuBi2O4 requires the absorption of higher energy photons that brings about the O 2p to Cu 3d LMCT transition (or O 2p to Bi 6p at even higher energies) [160].
Due to the nature of the optical transitions, the photogenerated electrons and holes are confined in the [CuO4]6‒ units that are isolated from each other, as mentioned above (Figure 8b). Consequently, there is a lack of DOS overlap among [CuO4]6‒ units that would serve as conduction paths for the minority and majority carriers. This results in a localization of electrons at discrete Cu sites, and hence, to polaron formation that is stabilized in part because of the attainment of a transient stable electronic configuration (from Cu2+ 3d9 to Cu+ 3d10). Consequently, electronic conduction in CuBi2O4 proceeds via small-polaron hopping between [CuO4]6‒ units along the c-axis, as the nearest cation to Cu is another Cu along the c-axis (Figure 8c) [156,158]. Similar to previous discussions, this small-polaron hopping conduction mechanism is responsible for the poor charge carrier transport in CuBi2O4. Based on its band gap of 1.5‒1.8 eV, the maximum theoretical photocurrent for CuBi2O4 under AM1.5 illumination is 19.7‒29.0 mA·cm−2 [164]. Actual photocurrents are therefore expected to be lowered by the inferior charge transport, and even more by the sluggish CO2 reduction interfacial kinetics [157]. Therefore, strategies to address the inherent limitations of CuBi2O4 are requisite for CuBi2O4 to be a suitable photocathode for PEC applications.
To achieve superior PEC performance, preparation of high quality CuBi2O4 photocathodes is of foremost importance. CuBi2O4 films have been prepared using simple techniques such as electrodeposition [162,166,169], drop-casting [156,170,171], and spin-coating [167,168]. However, these techniques usually resulted in incomplete surface coverage, which can reduce the photoactivity because exposed FTO can act as sites for back reactions [172,173]. In addition, nanoscale phase impurities that are potentially deleterious to the integrity of the photoelectrode and that may not be detected by XRD can also be formed [174]. More advanced techniques such as spray pyrolysis [173], reactive co-sputtering [157,175], and pulsed laser deposition [176,177] yielded highly phase-pure, homogeneous CuBi2O4 films. Furthermore, phase-purity was better achieved with rapid thermal processing (RTP) (10 min at 650°C) than conventional furnace heating (CFH) (72 h at 500°C) of pulsed-laser-deposition-prepared CuO/Bi2O3 precursor [178]. This finding was attributed to nucleation and grain growth processes during heating (Figure 9a). The high heating rate in RTP rapidly brings the CuO/Bi2O3 precursor to a high-temperature steady state that enables fast diffusion of Cu ions into the Bi2O3 layer. This yields a homogeneous distribution of CuBi2O4 nucleation sites after 5 min; rapid grain growth then results in the formation of a single-phase CuBi2O4. In contrast, the slower heating rate in CFH causes grain growth to occur even while Cu diffusion is still taking place. The resulting CuBi2O4 grains impede further diffusion so that phase transformation of the precursor is incomplete.
![Figure 9
(a) Comparison of the diffusion, nucleation, and grain growth processes during the formation of CuBi2O4 via rapid thermal processing and conventional furnace heating. (Reprinted from ref. [178]). (b) Scheme of preparation of nanodendritic CuBi2O4 (with TiO2 as protection layer). (Reprinted with permission from ref. [179]. Copyright © 2021, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_009.jpg)
(a) Comparison of the diffusion, nucleation, and grain growth processes during the formation of CuBi2O4 via rapid thermal processing and conventional furnace heating. (Reprinted from ref. [178]). (b) Scheme of preparation of nanodendritic CuBi2O4 (with TiO2 as protection layer). (Reprinted with permission from ref. [179]. Copyright © 2021, American Chemical Society).
Carrier diffusion length in CuBi2O4 was found to be in the order of 10‒60 nm, as determined from time-resolved microwave conductivity measurements [156]. Compared to the light penetration depth of 244 nm (for a wavelength of 550 nm), the short carrier diffusion length leads to excessive electron-hole recombination, severely limiting the PEC performance [156,164]. As with other photoelectrodes, the diffusion pathway of charge carriers can be shortened through nanostructuring, which may also increase internal light scattering and active sites for interfacial reactions [34,35,36,37,38,39,40]. Recently, a novel template-assisted synthesis strategy was devised to prepare a CuBi2O4 film with a nanodendritic structure, with the trunks and sub-branches having a radius around 90–150 nm (Figure 9b) [179]. The nanodendritic CuBi2O4 demonstrated a photocurrent (measured in N2-saturated 0.1 M Na2SO4 electrolyte under AM1.5 G illumination) twice that of a planar film, which the authors ascribed to a shorter diffusion length of the CuBi2O4/electrolyte interface as well as a higher specific surface area. To improve its photostability, the nanodendritic CuBi2O4 was conformally coated with a TiO2 protective overlayer.
Enhancing hole transport to minimize electron-hole recombination can significantly improve the performance of CuBi2O4 photocathodes. To this end, doping and use of hole transport layers have been investigated. Doping of CuBi2O4 with Ag resulted in an increase in hole concentration, as substitution of Ag+ with Bi3+ is charge-compensated by free holes (no Ag impurities were formed) [172]. Owing to the enhanced hole transport, Ag-doped CuBi2O4 exhibited higher photocurrent and better stability (against anodic photocorrosion) than undoped CuBi2O4 [172]. Hole transport layers (HTL) address the mismatch between the work function of FTO (a degenerately doped n-type semiconductor), the most commonly employed conducting substrate, and the Fermi level of the semiconductor photoelectrode. Due to such mismatch, a Schottky barrier may form at the FTO/CuBi2O4 interface, impeding hole collection in the back contact and enhancing carrier recombination [180]. Gold [162] and Cu-doped NiO [180] were found to be effective HTLs for CuBi2O4. The energy band alignment of the FTO/Cu:NiO/CuBi2O4 photocathode is shown in Figure 10a.
![Figure 10
(a) The favorable alignment of the energy levels of FTO/Cu:NiO/CuBi2O4 facilitates hole transport to FTO, while blocking electrons from reaching FTO. (Reprinted from ref. [180]). (b) Band alignment of CuBi2O4/Cu1.5TiO
z
interface, showing favorable electron transport to the overlayer. (Reprinted with permission from ref. [175]. Copyright © 2020, John Wiley and Sons). (c) Introduction of surface states on CuBi2O4 upon irradiation resulted in a shift in Fermi level and an increased band bending. (Reprinted from ref. [182]).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_010.jpg)
(a) The favorable alignment of the energy levels of FTO/Cu:NiO/CuBi2O4 facilitates hole transport to FTO, while blocking electrons from reaching FTO. (Reprinted from ref. [180]). (b) Band alignment of CuBi2O4/Cu1.5TiO z interface, showing favorable electron transport to the overlayer. (Reprinted with permission from ref. [175]. Copyright © 2020, John Wiley and Sons). (c) Introduction of surface states on CuBi2O4 upon irradiation resulted in a shift in Fermi level and an increased band bending. (Reprinted from ref. [182]).
Another strategy to improve the PEC cell performance of CuBi2O4 is the addition of conformal protective overlayers to address its photocorrosion. The function of the overlayer should not be limited to passivation of surface states and blocking of CuBi2O4/electrolyte contact; ideally, the overlayer should also act as an electron-selective contact that enhances charge separation [181,182]. Like in HTLs, addition of an overlayer will require a precise engineering of the interface such that energy levels are properly aligned and that no recombination traps are formed [175]. To this end, the multilayer scheme initially developed for Cu2O [104] was adopted for CuBi2O4 with CdS as the buffer layer between CuBi2O4 and the TiO2 protection layer [181]. In another study, high throughput methodology was employed to identify a suitable multinary metal oxide overlayer for CuBi2O4 [175]. An optimized Cu1.5TiO z overlayer was well-matched with CuBi2O4, forming a heterojunction that favored electron transport to the surface (Figure 10b). A unique, facile approach to introduce a protective overlayer on CuBi2O4 involves its pre-irradiation to form reduced Cu states on its surface [182]. The surface states induce an additional downward band bending upon illumination that contributes to improved charge transport and hence PEC cell performance (Figure 10c).
Tailoring the Cu/Bi ratio to deliberately introduce phase impurities has also been applied as a strategy to achieve better PEC cell performance of CuBi2O4. Similar to CuFeO2, nonstoichiometric CuBi2O4 films likely contain binary oxide phases, that is, CuO if Cu/Bi >1/2 or α,β-Bi2O3 otherwise, since CuBi2O4 is a line compound (there is no Cu x Bi2−x O4 solid solution) [174]. Because CuO and Bi2O3 are both photoelectrochemically active, the electronic and optical properties of these nonstoichiometric films can be tuned with the Cu/Bi ratio [163]. Since CuO has a narrower band gap (∼1.5 eV), the optical band gap redshifts and the films become progressively darker with higher Cu/Bi ratio [183]. Also, with higher Cu/Bi ratio, the flat band potential shifts to less positive values (away from valence band) because of a lower Cu vacancy concentration (recall that Cu vacancies introduce p-type behavior of CuBi2O4) [164,167,175]. Taking advantage of this concept, a photocathode with a Cu/Bi ratio gradient across its thickness was devised by placing into contact three films of varying stoichiometries (Cu/Bi = 1/3, 1/2, and 1) and thus, varying Fermi levels (Figure 11a) [164]. When placed in contact, the Fermi levels of the three films equilibrate by distributing the free holes, so that the conduction and valence bands bend at each interface (Figure 11b). The forward-gradient film, that is, with Cu/Bi ratio increasing from substrate side to electrolyte side (or Cu vacancy concentration decreasing), exhibited the correct band bending where photogenerated holes are transported to the FTO back contact, while the electrons are transported to the surface (Figure 11c), and resulted in the best photoresponse versus a homogeneous and a reverse-gradient film.
![Figure 11
(a) Band positions of nonstoichiometric CuBi2O4 films before contact. (b) Equilibration of Fermi levels and corresponding band bending at the interfaces upon contact. (c) Formation of CuBi2O4 with forward gradient. (Reprinted with permission from ref. [164]. Copyright © 2017, American Chemical Society).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_011.jpg)
(a) Band positions of nonstoichiometric CuBi2O4 films before contact. (b) Equilibration of Fermi levels and corresponding band bending at the interfaces upon contact. (c) Formation of CuBi2O4 with forward gradient. (Reprinted with permission from ref. [164]. Copyright © 2017, American Chemical Society).
Indeed, composites of CuBi2O4 with its secondary oxide CuO, both as randomly intermixed [163,168] or as separate layers [184,185], have demonstrated better PEC cell performance than CuBi2O4 or CuO alone. The simultaneous formation of both CuBi2O4 and CuO phases from a common precursor ensures a sufficient interfacial contact between the two phases, a requirement to establish an effective heterojunction [163,185]. However, the exact alignment of the heterojunction is still unclear. In one study, a layered CuO/CuBi2O4 was synthesized where CuBi2O4 acts as the electrocatalytic phase and as a protective layer for the relatively less stable CuO [184]. The authors claimed that a type II heterojunction forms at the interface, where photogenerated electrons are injected to CuBi2O4, while the holes to CuO (Figure 12a). In a different study, a CuBi2O4/CuO heterojunction with CuO deposited on CuBi2O4 was prepared [185]. CuO and CuBi2O4 still exhibited a type II alignment but with electron flow towards CuO (Figure 12b). Aside from CuO, composites of CuBi2O4 with Bi2O3 [163,186,187], as well as other oxides [188,189], have been reported in literature with varying successes in the improvement of PEC activity.
![Figure 12
Proposed energy band alignments of the CuBi2O4/CuO heterojunction: (a) electron moves from CuO to CuBi2O4 to electrolyte (Reprinted with permission from ref. [184]. Copyright © 2021, Elsevier) and (b) electron moves from CuBi2O4 to CuO to electrolyte (Reprinted from ref. [185]. Copyright © 2020, Royal Society of Chemistry).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_012.jpg)
Proposed energy band alignments of the CuBi2O4/CuO heterojunction: (a) electron moves from CuO to CuBi2O4 to electrolyte (Reprinted with permission from ref. [184]. Copyright © 2021, Elsevier) and (b) electron moves from CuBi2O4 to CuO to electrolyte (Reprinted from ref. [185]. Copyright © 2020, Royal Society of Chemistry).
Despite this significant progress with CuBi2O4, majority of the efforts were directed towards H2O reduction; studies on the improvement of CuBi2O4 as a photocathode for CO2 reduction are few. Electrodeposited CuBi2O4 films were studied for the PEC reduction of CO2 [166]. The film thickness was first optimized based on the tradeoff between an enhanced light absorption and an increase in number of recombination centers. The optimized sample was used as photocathode for CO2 reduction (in a CO2-saturated 0.1 M KHCO3 electrolyte, under a 300 W Xe lamp illumination and a bias of 0 V vs Ag/AgCl). CO was the only detected product in the headspace after 4 h of reaction, but with a very low yield (0.0155 μmol·cm−2). In another study, no product was even detected despite the conduction band being in a thermodynamically feasible position for CO2 reduction [190]. These results indicate that aside from the inherently poor charge transport kinetics, the bare CuBi2O4 surface shows little activity for CO2 reduction and that similar strategies to those discussed in the preceding paragraphs are necessary to drive CO2 reduction on CuBi2O4.
TiO2 was employed as a protective overlayer on nanoporous CuBi2O4, using two different overlayer thicknesses (50 nm and 1 μm) [191]. However, TiO2 was not conformally deposited on the CuBi2O4 surface; there were regions where there was contact between CuBi2O4 and the electrolyte. CO and H2 were the detected products in PEC measurements in a CO2-saturated bicarbonate electrolyte under an applied bias of −0.50 V (vs reversible hydrogen electrode (RHE)). Measured photocurrents and the overall yield were both larger for the sample with thicker TiO2 due to a more efficient charge separation versus the sample with thinner TiO2, where the space-charge region in the TiO2/electrolyte interface cannot be fully supported by the thin overlayer. However, selectivity for CO was lower for the sample with thicker TiO2 layer; the authors attributed this to the effect of the nanostructure on mass transport (Figure 13). In the sample with thicker TiO2 layer, the channels or pores were narrower so that local pH changes near the interface were more pronounced. Because CO2 reduction is more favored at higher pH, CO2 is more quickly exhausted near the interface, and the process becomes mass transfer limited. Consequently, the electrons are instead consumed for the HER.
![Figure 13
Effect of mesostructure on activity and selectivity of CuBi2O4/TiO2 nanocomposite for PEC reduction of CO2. (Reprinted with permission from ref. [191]. Copyright © 2021, Elsevier).](/document/doi/10.1515/rams-2022-0043/asset/graphic/j_rams-2022-0043_fig_013.jpg)
Effect of mesostructure on activity and selectivity of CuBi2O4/TiO2 nanocomposite for PEC reduction of CO2. (Reprinted with permission from ref. [191]. Copyright © 2021, Elsevier).
3.3 Other copper ternary oxides
Other copper delafossites, namely, CuCrO2 and CuGaO2, have been investigated as photocathodes for CO2 reduction. CuCrO2 is, however, a wide band gap p-type semiconductor (2.95‒3.15 eV) implying that UV illumination is necessary to obtain a photoresponse [192]. CuCrO2 prepared via solution combustion synthesis was tested for its photoactivity towards CO2 reduction [193]. The as-synthesized sample contained phase impurities, namely, CuO (4 wt%) and α-Cr2O3 (6 wt%); annealing in Ar at 700°C allowed the formation of a more phase-pure sample. As expected, onset of IPCE was near the UV region (438 nm). Long-term photoelectrolysis of the Ar-annealed sample in CO2-saturated bicarbonate, under UV illumination, resulted in the formation of CO, CH4, and H2 in the gas phase, while HCOOH and CH3OH in the liquid phase, with a total Faradaic efficiency of 85‒90%. Although photocurrents decreased significantly after one hour, the authors claimed that their CuCrO2 photocathode was relatively more stable than Cu2O. Because unmodified CuCrO2 was proven capable of CO2 reduction, there is much room for improvement especially by making it responsive to visible light. CuGaO2 is another wide band gap p-type semiconductor (3.3‒3.75 eV) which has been employed as hole transport layer in dye-sensitized solar cells [194,195]. The photochemical activity (in powder form) of CuGaO2 and its solid solution with CuFeO2 and CuGa1‒x Fe x O2 towards CO2 reduction was investigated [194]. The indirect band gap of 2.55 eV in pure CuGaO2 (which was not previously observed in thin film CuGaO2) was observed to shift to 1.5 eV upon Fe alloying, due to a crystal strain that modifies the electronic structure of CuGaO2. Observed reduction products from an aqueous CO2 solution upon illumination with broadband light were CO and CH4 (trace). However, the amount of CO evolved did not correlate with the degree of Fe substitution, so it is likely that the indirect band gap absorption does not result in an efficient charge separation and accordingly contribute to the photocatalytic process. In a different study, CuGaO2 was employed as a scaffold for a Ru(ii)‒Re(i) supramolecular photocatalyst for CO2 reduction [196]. PEC testing using the hybrid photocathode in a CO2-saturated aqueous solution under visible light irradiation (>460 nm) produced both CO and H2, with a total Faradaic efficiency of 81%. A PEC cell employing the Ru(ii)‒Re(i)/CuGaO2 hybrid photocathode coupled to a CoO x /TaON photoanode with a Z-scheme configuration was then devised. Under visible light illumination and bias-free condition, CO and H2 were produced at the cathode, while O2 evolution was observed at the anode.
Other Cu2+-based ternary oxides that have been investigated as photocathodes for CO2 reduction are CuFe2O4, Cu2V2O7, and the copper niobates CuNb2O6 and Cu3Nb2O8. CuFe2O4 is a ternary oxide with an inverse spinel structure and a band gap of 1.39‒1.42 eV. Methanol was the only product detected in the liquid phase from PEC reduction in a CO2-saturated bicarbonate solution under visible light irradiation (>470 nm), with a corresponding Faradaic efficiency of 62% [197]. Addition of CdS [198], graphene oxide [199], and polyaniline (PANI) [200] as co-catalysts resulted in a better PEC performance, as reflected in the higher IPCEs of 12.09, 8.02, and 7.1% (at 470 nm), respectively, (compared to 5.1% for bare CuFe2O4) as well as a higher selectivity for methanol formation, 72, 87, and 73%, respectively (compared to 62% for bare CuFe2O4). This improvement was attributed to an enhanced electron-hole separation and to an increased CO2 adsorption for the case of PANI. A direct Z-scheme system of Cu2V2O7 (with a band gap of 2.01 eV) and g-C3N4 was devised [201]. Under illumination of light with wavelength of 400‒700 nm, the Cu2V2O7/g-C3N4 system produced a much higher photocurrent response than Cu2V2O7 or g-C3N4 alone. CO2 photoreduction (under a photocatalytic scheme) resulted in the formation of CH4, CO, and O2 with rates of 305, 166, and 706 μmol·g‒1 catalyst·h‒1, respectively. CuNb2O6, belonging to a class of metal niobates with an orthorhombic columbite structure, was found to be photoactive for PEC CO2 reduction even without any co-catalyst [202]. It showed photoresponse under visible light illumination (150 W tungsten-halogen lamp), consistent with its optical band gap of 1.77 eV. Photocurrents measured in CO2-saturated 0.1 M NaHCO3 solution (pH 7) were up to six times greater than those measured in an N2-saturated 0.1 M Na2SO4 solution (pH 7), indicating that the photocurrent was being used in the reduction of CO2, although no efforts were done to identify and quantify the reduction products. Another copper niobate, Cu3Nb2O8 (with band gap of 2.5 eV), was also found to photoelectrochemically reduce CO2 under AM 1.5 G illumination, with CO as the primary product [203]. However, the Faradaic efficiency for CO evolution was found to be around 9% (at ‒0.20 V vs Ag/AgCl), indicating that majority of the cathodic photocurrent was being used in the photocorrosion of the Cu3Nb2O8 film. Indeed, chronoamperometry tests (in CO2-saturated 0.5 M NaHCO3 solution at ‒0.20 V vs Ag/AgCl) showed a 98% decrease in the initial photocurrent only after 20 min. Reannealing of the tested films restored the initial photocurrent to that of the as-synthesized film, corroborating the cathodic photocorrosion of Cu3Nb2O8, and the use of protective layers or addition of co-catalysts will be needed to improve its PEC performance.
The results of PEC CO2 reduction studies on the copper ternary oxides, where the reduction products were identified and quantified, are summarized in Table 3. Although beyond the scope of this review, there are many research efforts in recent years to explore other classes of materials as photocathodes. Non-copper ternary oxides, such as AgRhO2 [204], BiFeO3 [205,206], CaFe2O4 [207,208], Ca2Fe2O5 [209,210], LaFeO3 [211,212], NiFeO4 [213], and PbMoO4 [214], have been prepared as photocathodes but their application towards PEC CO2 reduction is yet to be investigated. On the other hand, copper-based chalcogenides, such as CuInS2 [215,216,217,218], CuGaS2 [219], CuGa3Se5 [220], Cu(In,Ga)Se2 [221,222], Cu2ZnSnS4 [223,224], and Cu2ZnGeS4 [225], which have been widely used in solar cells due to their favorable transport properties, have recently been investigated as CO2 reduction photocathodes (Table 4). Although their efficiency and stability are comparable with the copper ternary oxides, the cost and toxicity (primarily of In and Ga), as well as the complicated preparation methods that limit possibilities for scale-up, are major issues that need to be addressed [226].
Ternary copper oxide photocathodes for PEC CO2 reduction
Photocathode | Reaction conditions | Products formed | Photostability | Ref. |
---|---|---|---|---|
(a) CuFeO 2 | ||||
Mg-doped CuFeO2 | Blue LED illumination (470 nm, 2.1 mW·cm−2), CO2-saturated 0.1 M NaHCO3 | HCOO− (FE: 10%) at −0.9 V vs SCE, 8–24 h | — | [123] |
CuFeO2/CuO | 150 W Xe arc lamp, CO2-saturated 0.1 M KHCO3 | HCOO− (∼1% solar-to-formate efficiency, 90% selectivity) using a two-electrode system (no external bias) | Formate production continued over 17 days, but rate is reduced by 60%; reannealing of the photoelectrode restored its activity | [148,150] |
CuFeO2/CuO | Warm white light LED (100 mW·cm−2), CO2-saturated 0.1 M KHCO3 | CH3COO− (FE: 80%) at −0.4 V vs Ag/AgCl, 2 h | Acetate formation quickly deactivates after first 10 min of testing due to dissolution of Fe | [149] |
CuFeO2/Cu2O nanorods | 300 W Xe arc lamp, CO2-saturated 0.5 M KHCO3 | CH3COO− (FE: 68.6%), HCOO−(21.6%) at 0.35 V vs RHE, 30 min | Photocurrent decreased by 8% after 10 LSV repetitions | [151] |
CuFeO2/Nb-doped TiO2 nanotubes | 250 W Xe arc lamp, CO2-saturated 0.1 M NaHCO3 | C2H5OH, CH2O at −0.4 V vs SCE, 5 h | — | [147] |
(b) CuBi 2 O 4 | ||||
CuBi2O4 | 300 W Xe lamp, CO2-saturated 0.1 M KHCO3 | CO (0.0155 μmol·cm−2) at 0 V vs Ag/AgCl, 4 h | — | [166] |
Au-decorated CuBi2O4 | 300 W Xe lamp, CO2-saturated 0.1 M KHCO3 | No product detected | — | [190] |
CuBi2O4/TiO2 | 300 W Xe lamp, CO2-saturated 0.1 M KHCO3 | CO at −0.5 V vs RHE | — | [191] |
(c) Other Cu ternary oxides | ||||
CuCrO2 | 300 W Hg-Xe arc lamp, CO2-saturated 0.1 M NaHCO3 | CO (5.9%), HCOOH (2.7%), CH3OH (0.5%), CH4 (0.7%) at 0 V vs RHE, 4 h | Photocurrent decreased substantially after 1 h of testing | [193] |
CuGaO2/Ru(ii)-Re(i) | 300 W Xe lamp, CO2-saturated 0.05 M NaHCO3 | CO (FE: 72% for total cathodic reaction including H2 evolution) in a non-biased cell with CoO x /TaON as photoanode | About 80% of the Ru(ii)-Re(i) catalyst lost electrochemical activity after 15 h | [196] |
CuFe2O4 | XD 300 high brightness cold light, CO2-saturated 0.1 M NaHCO3 | CH3OH (FE: 87% for CuFe2O4/GO at −0.4 V vs Ag/AgCl, 72% for CuFe2O4/CdS at −0.35 V vs Ag/AgCl, 73% for CuFe2O4/PANI at −0.4 V vs Ag/AgCl) | For CuFe2O4/GO, little decrease in CO2 reduction yield after 4 cycles of testing | [198,199,200] |
Cu3Nb2O8 | Solar simulator (100 mW·cm−2), CO2-saturated 0.5 M NaHCO3 | CO (FE: 9%) at −0.2 V vs Ag/AgCl, 20 min | Rapid decrease in photocurrent after 800 s due to reduction of Cu; reannealing restores initial photocurrent | [203] |
FE, Faradaic efficiency.
Copper chalcogenide photocathodes for PEC CO2 reduction
Photocathode | Reaction conditions | Products formed | Photostability | Ref |
---|---|---|---|---|
CuInS2/graphene | Xe lamp (100 mW·cm−2), 0.1 M acetate buffer | CH3OH (95%) at −0.59 V vs SCE, 5 h | No change in morphology and structure was detected after 10 h testing | [215] |
CuInS2/CuO/CuS | Xe lamp (100 mW·cm−2), 0.1 M acetate buffer with 10 mM pyridine | CH3OH (86.8%) at −0.60 V vs SCE, 1.5 h | No significant change in the crystalline structure after 4.5 h testing, but particle size became smaller | [216] |
CuInS2/Cu-In | Xe lamp (100 mW·cm−2), 0.1 M KHCO3 | C2H5OH (FE: ∼90%), CH3OH (trace) at −0.70 V vs SCE, 1.5 h | — | [217] |
CuInS2/CuFeO2 | Xe lamp (100 mW·cm−2), 0.1 M acetate buffer with 10 mM pyridine | CH3OH (88%) at −0.70 V vs SCE, 1.5 h | Rate of methanol formation remains almost constant within 9 h and then decreases | [218] |
CuGaS2/CdS/TiO2 | 150 W solar simulator lamp, 0.1 M Na2SO4 | CH3OH (65 mol%), C2H5OH (22.5%), (CH3)2CO (12.5%) at −0.70 V vs Ag/AgCl, 4 h | CH3OH, C2H5OH, and (CH3)2CO productions dropped by 12%, 56%, and 60% in 2nd use of photocathode | [219] |
CuGa3Se5/CdS/N,N′-(1,4-phenylene)bispyridinium ditriflate | Monochromatic LED (450 nm, 65 mW·cm−2), 0.1 M KHCO3 | CO at 0 V vs RHE, 2 h | The molecular additive resulted in lower coarsening and improved retention of the CdS coating | [220] |
CIGS/TiO2/Co-qPyH | 150 W Xe lamp, 0.1 M KHCO3 | CO (97% selectivity) at −0.06 V vs RHE, 2 h | Photocurrent showed only a slight decay after 2 h electrolysis | [221] |
CIGS/CdS/ZnO/AZO | 300 W Xe arc lamp, 0.5 M KHCO3 | CO (FE: 99.3%) at −0.2 V vs RHE; CO (49.3%), HCOOH (15.4%), CH3OH (8.45%) at −0.5 V vs RHE | — | [222] |
CZTS/[RuCE + RuCA] | Xe light source, purified water | HCOO− (FE: 82%) at −0.4 V vs RHE, 3 h | — | [223] |
CZTS/CdS | Solar simulator (100 mW·cm−2), 0.1 M KHCO3 | CO, CH3OH, C2H5OH (measured at −0.4, −0.5, −0.6, −0.7, and −0.8 V vs RHE) | — | [224] |
Zn0.5Cu2ZnGeS4 | Solar simulator, 0.1 M KHCO3 | CO (FE: 3.3%) at −0.2 V vs RHE, 2 h | — | [225] |
FE, Faradaic efficiency.
4 Summary and outlook
In the search for new p-type materials suitable as photocathode for PEC applications, copper ternary oxides have recently gained attention due to their favorable band structure and earth-abundant nature. This article examined the current progress achieved with CuFeO2 and CuBi2O4 by summarizing the various strategies to improve their efficiency and stability towards CO2 reduction. Other copper ternary oxides, CuCrO2, CuGaO2, CuFe2O4, Cu2V2O7, CuNb2O6, and Cu3Nb2O8, were also reviewed. For these photocathodes, CO, HCOOH, HCOH, CH3OH, and CH4 as well as C2H5OH and (CH3)2CO were the observed CO2 reduction products (Table 3). However, photocorrosion is still significant in these materials. Several studies are in search of selective inhibitors that can allow only specific reactions to proceed, aiding the overall kinetics. Polymers and membranes have been used recently as the protective overlayer that simultaneously fosters CO2 binding sites [227] and modifies the surface hydrophobicity to improve selectivity [228].
Most of the papers reviewed here approach the improvement in the copper oxide photocathodes from the viewpoint of charge transport. In Cu2O, it has been observed that surface recombination and bulk transport were the contributors to the lifetime of charges [229]. A close connection between semiconductor and conducting support leads to faster charge separation; the challenge is to determine the type and structure of the support and the copper oxides. However, a system of efficient and stable CO2 reduction relies not only on charge transport but also on chemical reaction kinetics and mass transfer of reactants and products. Theoretical studies on CO2 adsorption and reaction mechanisms are being done by several groups [230,231]. Similarly, there have also been theoretical and experimental works to understand CO2 reduction active sites in copper and copper-based semiconductors [228,232]. Studies such as these can guide synthesis of structures with complementary adsorption and active site.
It remains a prospect whether a semiconductor material with the correct compositional mix that satisfies all requirements as photoelectrode can be identified. More likely, this semiconductor will come from complex metal oxides where there are greater possibilities for tuning and optimization of material properties. With this large material space, theory on interfacial phenomena, such as CO2 activation and intermediates, should be refined much further to establish advanced descriptors for computational screening of new complex oxides. High throughput methodologies for the synthesis of a large library of candidate materials and testing of their relevant properties for PEC applications will further accelerate materials discovery. It must be noted, however, that with increasing complexity of multinary oxides, the cost and ease of processability, along with sustainability considerations critical towards large scale production, become more difficult to attain.
In the past few decades, significant achievements have been made in the field of PEC solar fuel generation. At its current state, further advances will rely not only on materials development, but also on device modeling and simulation and techno-economic analysis. For PEC CO2 reduction to become a viable, carbon-neutral technology that can harness the enormous, undeveloped potential of solar energy, many key challenges remain to be solved.
-
Funding information: This work was supported by the Engineering Research and Development for Technology (ERDT) of the Department of Science and Technology (DOST).
-
Author contributions: I.L.E. Gonzaga: conceptualization and writing – original draft, review & editing; C.C. Mercado: conceptualization, writing – review & editing, and supervision.
-
Conflict of interest: Authors state no conflict of interest.
References
[1] Lewis, N. and D. Nocera. Powering the planet: Chemical challenges in solar energy utilization. Proceedings of the National Academy of Sciences, Vol. 103, No. 43, 2006, pp. 15729–15735.10.1073/pnas.0603395103Search in Google Scholar
[2] Pleskov, Y. Solar energy conversion in photoelectrochemical cells with semiconductor electrodes. Progress in Surface Science, Vol. 15, No. 4, 1984, pp. 401–456.10.1016/0079-6816(84)90005-4Search in Google Scholar
[3] Lewis, N. Research opportunities to advance solar energy utilization. Science, Vol. 351, No. 6271, 2016, id. aad1920.10.1126/science.aad1920Search in Google Scholar PubMed
[4] Centi, G. and S. Perathoner. Towards solar fuels from water and CO2. Chemsuschem, Vol. 3, No. 2, 2010, pp. 195–208.10.1002/cssc.200900289Search in Google Scholar PubMed
[5] Yan, Y., J. Gu, E. L. Zeitler, and A. B. Bocarsly. Photoelectrocatalytic reduction of carbon dioxide. In Carbon dioxide utilization: Closing the carbon cycle, 1st edn, Styring, P., A. Quadrelli, and K. Armstrong, Eds, Elsevier Inc., Amsterdam, The Netherlands, 2015, pp. 211–233.10.1016/B978-0-444-62746-9.00012-8Search in Google Scholar
[6] Tilley, S. D. Recent advances and emerging trends in photo‐electrochemical solar energy conversion. Advanced Energy Materials, Vol. 9, No. 2, 2019, id. 1802877.10.1002/aenm.201802877Search in Google Scholar
[7] Gust, D., T. Moore, and A. Moore. Solar fuels via artificial photosynthesis. Accounts of Chemical Research, Vol. 42, No. 12, 2009, pp. 1890–1898.10.1021/ar900209bSearch in Google Scholar PubMed
[8] Roy, S., O. Varghese, M. Paulose, and C. Grimes. Toward solar fuels: Photocatalytic conversion of carbon dioxide to hydrocarbons. ACS Nano, Vol. 4, No. 3, 2010, pp. 1259–1278.10.1021/nn9015423Search in Google Scholar PubMed
[9] Hoffmann, M., J. Moss, and M. Baum. Artificial photosynthesis: semiconductor photocatalytic fixation of CO2 to afford higher organic compounds. Dalton Transactions, Vol. 40, No. 19, 2011, pp. 5151–5158.10.1039/c0dt01777aSearch in Google Scholar PubMed
[10] Barber, J. and P. Tran. From natural to artificial photosynthesis. Journal of The Royal Society Interface, Vol. 10, No. 81, 2013, id. 20120984.10.1098/rsif.2012.0984Search in Google Scholar PubMed PubMed Central
[11] Krol, R. and M. Grätzel. Photoelectrochemical hydrogen production. Springer, New York, NY, 2012.10.1007/978-1-4614-1380-6Search in Google Scholar
[12] Olah, G., G. Prakash, and A. Goeppert. Anthropogenic chemical carbon cycle for a sustainable future. Journal of The American Chemical Society, Vol. 133, No. 33, 2011, pp. 12881–12898.10.1021/ja202642ySearch in Google Scholar
[13] Spinner, N., J. Vega, and W. Mustain. Recent progress in the electrochemical conversion and utilization of CO2. Catalysis Science & Technology, Vol. 2, No. 1, 2012, pp. 19–28.10.1039/C1CY00314CSearch in Google Scholar
[14] White, J., M. Baruch, J. Pander, Y. Hu, I. Fortmeyer, J. Park, et al. Light-driven heterogeneous reduction of carbon dioxide: photocatalysts and photoelectrodes. Chemical Reviews, Vol. 115, No. 23, 2015, pp. 12888–12935.10.1021/acs.chemrev.5b00370Search in Google Scholar
[15] Montoya, J., L. Seitz, P. Chakthranont, A. Vojvodic, T. Jaramillo, and J. Nørskov. Materials for solar fuels and chemicals. Nature Materials, Vol. 16, No. 1, 2016, pp. 70–81.10.1038/nmat4778Search in Google Scholar
[16] Kumaravel, V., J. Bartlett, and S. Pillai. Photoelectrochemical conversion of carbon dioxide (CO2) into fuels and value-added products. ACS Energy Letters, Vol. 5, No. 2, 2020, pp. 486–519.10.1021/acsenergylett.9b02585Search in Google Scholar
[17] Mikkelsen, M., M. Jørgensen, and F. C. Krebs. The teraton challenge. A review of fixation and transformation of carbon dioxide. Energy & Environmental Science, Vol. 3, No. 1, 2010, pp. 43–81.10.1039/B912904ASearch in Google Scholar
[18] Greenblatt, J., D. Miller, J. Ager, F. Houle, and J. Sharp. The technical and energetic challenges of separating (photo)electrochemical carbon dioxide reduction products. Joule, Vol. 2, No. 3, 2018, pp. 381–420.10.1016/j.joule.2018.01.014Search in Google Scholar
[19] Freund, H-J. and M. W. Roberts. Surface chemistry of carbon dioxide. Surface Science Reports, Vol. 25, No. 8, 1996, pp. 225–273.10.1016/S0167-5729(96)00007-6Search in Google Scholar
[20] Kumar, B., M. Llorente, J. Froehlich, T. Dang, A. Sathrum, and C. P. Kubiak. Photochemical and photoelectrochemical reduction of CO2. Annual Review of Physical Chemistry, Vol. 63, 2012, pp. 541–569.10.1146/annurev-physchem-032511-143759Search in Google Scholar PubMed
[21] Aresta, M. and A. Angelini. The carbon dioxide molecule and the effects of its interaction with electrophiles and nucleophiles. In Carbon dioxide and organometallics, Lu X. B., ed., Springer, Cham, 2016, pp. 1–38.10.1007/3418_2015_93Search in Google Scholar
[22] Costentin, C., M. Robert, and J-M. Savéant. Catalysis of the electrochemical reduction of carbon dioxide. Chemical Society Reviews, Vol. 42, No. 6, 2013, pp. 2423–2436.10.1039/C2CS35360ASearch in Google Scholar PubMed
[23] Dodson, L. G., M. C. Thompson, and J. M. Weber. Characterization of intermediate oxidation states in CO2 activation. Annual Review of Physical Chemistry, Vol. 69, 2018, pp. 231–252.10.1146/annurev-physchem-050317-021122Search in Google Scholar PubMed
[24] Álvarez, A., M. Borges, J. J. Corral‐Pérez, J. G. Olcina, L. Hu, D. Cornu, et al. CO2 activation over catalytic surfaces. ChemPhysChem, Vol. 18, No. 22, 2017, pp. 3135–3141.10.1002/cphc.201700782Search in Google Scholar PubMed
[25] Habisreutinger, S., L. Schmidt-Mende, and J. Stolarczyk. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angewandte Chemie International Edition, Vol. 52, No. 29, 2013, pp. 7372–7408.10.1002/anie.201207199Search in Google Scholar PubMed
[26] Peng, C., G. Reid, H. Wang, and P. Hu. Perspective: photocatalytic reduction of CO2 to solar fuels over semiconductors. The Journal of Chemical Physics, Vol. 147, No. 3, 2017, id. 030901.10.1063/1.4985624Search in Google Scholar PubMed
[27] Stolarczyk, J. K., S. Bhattacharyya, L. Polavarapu, and J. Feldmann. Challenges and prospects in solar water splitting and CO2 reduction with inorganic and hybrid nanostructures. ACS Catalysis, Vol. 8, No. 4, 2018, pp. 3602–3635.10.1021/acscatal.8b00791Search in Google Scholar
[28] Du, C., X. Wang, W. Chen, S. Feng, J. Wen, and Y. A. Wu. CO2 transformation to multicarbon products by photocatalysis and electrocatalysis. Materials Today Advances, Vol. 6, 2020, id. 100071.10.1016/j.mtadv.2020.100071Search in Google Scholar
[29] Liu, Y. and L. Guo. On factors limiting the performance of photoelectrochemical CO2 reduction. The Journal of Chemical Physics, Vol. 152, No. 10, 2020, id. 100901.10.1063/1.5141390Search in Google Scholar PubMed
[30] Butler, M. A. and D. S. Ginley. Principles of photoelectrochemical, solar energy conversion. Journal of Materials Science, Vol. 15, No. 1, 1980, pp. 1–19.10.1007/BF00552421Search in Google Scholar
[31] Lewis, N. S. and G. A. Shreve. Photochemical and photoelectrochemical reduction of carbon dioxide. In Electrochemical and electrocatalytic reactions of carbon dioxide, Sullivan, B. P., K. Krist, and H. E. Guard, Eds, Elsevier Inc., Amsterdam, The Netherlands, 1993, pp. 263–289.10.1016/B978-0-444-88316-2.50012-7Search in Google Scholar
[32] Cole, E. B. and A. B. Bocarsly. Photochemical, electrochemical, and photoelectrochemical reduction of carbon dioxide. In Carbon dioxide as chemical feedstock, M. Aresta, Ed, John Wiley & Sons, Ltd, WILEY-VCH Verlag GmbH & Co., KGaA, Weinheim, Germany, 2010, pp. 291–316.10.1002/9783527629916.ch11Search in Google Scholar
[33] Guijarro, N., M. S. Prévot, and K. Sivula. Surface modification of semiconductor photoelectrodes. Physical Chemistry Chemical Physics, Vol. 17, No. 24, 2015, pp. 15655–15674.10.1039/C5CP01992CSearch in Google Scholar
[34] Li, J. and N. Wu. Semiconductor-based photocatalysts and photoelectrochemical cells for solar fuel generation: a review. Catalysis Science & Technology, Vol. 5, No. 3, 2015, pp. 1360–1384.10.1039/C4CY00974FSearch in Google Scholar
[35] Sivula, K. and R. van de Krol. Semiconducting materials for photoelectrochemical energy conversion. Nature Reviews Materials, Vol. 1, No. 2, 2016, pp. 1–16.10.1038/natrevmats.2015.10Search in Google Scholar
[36] Tu, W., Y. Zhou, and Z. Zou. Photocatalytic conversion of CO2 into renewable hydrocarbon fuels: state-of-the-art accomplishment, challenges, and prospects. Advanced Materials, Vol. 26, No. 27, 2014, pp. 4607–4626.10.1002/adma.201400087Search in Google Scholar PubMed
[37] Marszewski, M., S. Cao, J. Yu, and M. Jaroniec. Semiconductor-based photocatalytic CO2 conversion. Materials Horizons, Vol. 2, No. 3, 2015, pp. 261–278.10.1039/C4MH00176ASearch in Google Scholar
[38] Xie, S., Q. Zhang, G. Liu, and Y. Wang. Photocatalytic and photoelectrocatalytic reduction of CO2 using heterogeneous catalysts with controlled nanostructures. Chemical Communications, Vol. 52, No. 1, 2016, pp. 35–59.10.1039/C5CC07613GSearch in Google Scholar PubMed
[39] Vu, N., S. Kaliaguine, and T. Do. Critical aspects and recent advances in structural engineering of photocatalysts for sunlight‐driven photocatalytic reduction of CO2 into fuels. Advanced Functional Materials, Vol. 29, No. 31, 2019, id. 1901825.10.1002/adfm.201901825Search in Google Scholar
[40] Ding, P., T. Jiang, N. Han, and Y. Li. Photocathode engineering for efficient photoelectrochemical CO2 reduction. Materials Today Nano, Vol. 10, 2020, id. 100077.10.1016/j.mtnano.2020.100077Search in Google Scholar
[41] Chen, P., Y. Zhang, Y. Zhou, and F. Dong. Photoelectrocatalytic carbon dioxide reduction: fundamental, advances and challenges. Nano Materials Science, Vol. 3, No. 4, 2021, pp. 344–367.10.1016/j.nanoms.2021.05.003Search in Google Scholar
[42] Rajeshwar, K., P. A. Maggard, and S. O’Donnell. In search of the “perfect” inorganic semiconductor/liquid interface for solar water splitting. The Electrochemical Society Interface, Vol. 30, No. 1, 2021, pp. 47–51.10.1149/2.F07211IFSearch in Google Scholar
[43] Wick, R. and D. D. Tilley. Photovoltaic and photoelectrochemical solar energy conversion with Cu2O. The Journal of Physical Chemistry C, Vol. 119, No. 47, 2015, pp. 26243–26257.10.1021/acs.jpcc.5b08397Search in Google Scholar
[44] Moakhar, R. S., S. M. Hosseini‐Hosseinabad, S. Masudy‐Panah, A. Seza, M. Jalali, H. Fallah‐Arani, et al. Photoelectrochemical water‐splitting using CuO‐based electrodes for hydrogen production: a review. Advanced Materials, Vol. 33, No. 33, 2021, id. 2007285.10.1002/adma.202007285Search in Google Scholar PubMed
[45] Sullivan, I., B. Zoellner, and P. A. Maggard. Copper(I)-based p-type oxides for photoelectrochemical and photovoltaic solar energy conversion. Chemistry of Materials, Vol. 28, No. 17, 2016, pp. 5999–6016.10.1021/acs.chemmater.6b00926Search in Google Scholar
[46] Lumley, M. A., A. Radmilovic, Y. J. Jang, A. E. Lindberg, and K. S. Choi. Perspectives on the development of oxide-based photocathodes for solar fuel production. Journal of The American Chemical Society, Vol. 141, No. 46, 2019, pp. 18358–18369.10.1021/jacs.9b07976Search in Google Scholar PubMed
[47] Li, C., J. He, Y. Xiao, Y. Li, and J-J. Delaunay. Earth-abundant Cu-based metal oxide photocathodes for photoelectrochemical water splitting. Energy & Environmental Science, Vol. 13, No. 10, 2020, pp. 3269–3306.10.1039/D0EE02397CSearch in Google Scholar
[48] Wang, K., Y. Ma, Y. Liu, W. Qiu, Q. Wang, X. Yang, et al. Insights into the development of Cu-based photocathodes for carbon dioxide (CO2) conversion. Green Chemistry, Vol. 23, No. 9, 2021, pp. 3207–3240.10.1039/D0GC04417BSearch in Google Scholar
[49] Abdi, F. F. and S. P. Berglund. Recent developments in complex metal oxide photoelectrodes. Journal of Physics D: Applied Physics, Vol. 50, No. 19, 2017, id. 193002.10.1088/1361-6463/aa6738Search in Google Scholar
[50] Rajeshwar, K., M. K. Hossain, R. T. Macaluso, C. Janáky, A. Varga, and P. J. Kulesza. Review – Copper oxide-based ternary and quaternary oxides: where solid-state chemistry meets photoelectrochemistry. Journal of The Electrochemical Society, Vol. 165, No. 4, 2018, pp. H3192–H3206.10.1149/2.0271804jesSearch in Google Scholar
[51] He, H., A. Liao, W. Guo, W. Luo, Y. Zhou, and Z. Zou. State-of-the-art progress in the use of ternary metal oxides as photoelectrode materials for water splitting and organic synthesis. Nano Today, Vol. 28, 2019, id. 100763.10.1016/j.nantod.2019.100763Search in Google Scholar
[52] Gerischer, H. Electrochemical behavior of semiconductors under illumination. Journal of The Electrochemical Society, Vol. 113, No. 11, 1966, id. 1174.10.1149/1.2423779Search in Google Scholar
[53] Bard, A. Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors. Journal of Photochemistry, Vol. 10, No. 1, 1979, pp. 59–75.10.1016/0047-2670(79)80037-4Search in Google Scholar
[54] Memming, R. Solar energy conversion by photoelectrochemical processes. Electrochimica Acta, Vol. 25, No. 1, 1980, pp. 77–88.10.1016/0013-4686(80)80054-5Search in Google Scholar
[55] Chang, X., T. Wang, P. Yang, G. Zhang, and J. Gong. The development of cocatalysts for photoelectrochemical CO2 reduction. Advanced Materials, Vol. 31, No. 31, 2019, id. 1804710.10.1002/adma.201804710Search in Google Scholar
[56] Gurevich, Y. Y. and Y. V. Pleskov. Photoelectrochemistry of semiconductors. Elsevier – Semiconductors and Semimetals, Vol. 19, 1983, pp. 255–328.10.1016/S0080-8784(08)60277-XSearch in Google Scholar
[57] Nozik, A. J. and R. Memming. Physical chemistry of semiconductor − liquid interfaces. The Journal of Physical Chemistry, Vol. 100, No. 31, 1996, pp. 13061–13078.10.1021/jp953720eSearch in Google Scholar
[58] Boston, D., K. L. Huang, N. de Tacconi, N. Myung, F. Macdonell, and K. Rajeshwar. Electro-and photocatalytic reduction of CO2: The homogeneous and heterogeneous worlds collide? In Photoelectrochemical water splitting: materials, processes and architectures, Lewerenz, H-J. and L. Peter, Eds, The Royal Society of Chemistry, Cambridge, UK, 2013, pp. 289–332.10.1039/9781849737739-00289Search in Google Scholar
[59] Rajeshwar, K., A. Thomas, and C. Janáky. Photocatalytic activity of inorganic semiconductor surfaces: myths, hype, and reality. The Journal of Physical Chemistry Letters, Vol. 6, No. 1, 2015, pp. 139–147.10.1021/jz502586pSearch in Google Scholar PubMed
[60] Barton, E., D. Rampulla, and A. Bocarsly. Selective solar-driven reduction of CO2 to methanol using a catalyzed p-GaP based photoelectrochemical cell. Journal of The American Chemical Society, Vol. 130, No. 20, 2008, pp. 6342–6344.10.1021/ja0776327Search in Google Scholar PubMed
[61] Ghadimkhani, G., N. R. de Tacconi, W. Chanmanee, C. Janaky, and K. Rajeshwar. Efficient solar photoelectrosynthesis of methanol from carbon dioxide using hybrid CuO–Cu2O semiconductor nanorod arrays. Chemical Communications, Vol. 49, No. 13, 2013, pp. 1297–1299.10.1039/c2cc38068dSearch in Google Scholar
[62] Halmann, M. Photoelectrochemical reduction of aqueous carbon dioxide on p-type gallium phosphide in liquid junction solar cells. Nature, Vol. 275, No. 5676, 1978, pp. 115–116.10.1038/275115a0Search in Google Scholar
[63] Aurian-Blajeni, B., M. Halmann, and J. Manassen. Electrochemical measurement on the photoelectrochemical reduction of aqueous carbon dioxide on p-Gallium phosphide and p-Gallium arsenide semiconductor electrodes. Solar Energy Materials, Vol. 8, No. 4, 1983, pp. 425–440.10.1016/0165-1633(83)90007-2Search in Google Scholar
[64] Ito, K., S. Ikeda, M. Yoshida, S. Ohta, and T. Iida. On the reduction products of carbon dioxide at a p-type gallium phosphide photocathode in aqueous electrolytes. Bulletin of the Chemical Society of Japan, Vol. 57, No. 2, 1984, pp. 583–584.10.1246/bcsj.57.583Search in Google Scholar
[65] Ikeda, S., A. Yamamoto, H. Noda, M. Maeda, and K. Ito. Influence of surface treatment of the p-GaP photocathode on the photoelectrochemical reduction of carbon dioxide. Bulletin of the Chemical Society of Japan, Vol. 66, No. 9, 1993, pp. 2473–2477.10.1246/bcsj.66.2473Search in Google Scholar
[66] Hirota, K., D. A. Tryk, T. Yamamoto, K. Hashimoto, M. Okawa, and A. Fujishima. Photoelectrochemical reduction of CO2 in a high-pressure CO2+ methanol medium at p-type semiconductor electrodes. The Journal of Physical Chemistry B, Vol. 102, No. 49, 1998, pp. 9834–9843.10.1021/jp9822945Search in Google Scholar
[67] Canfield, D. and K. W. Frese Jr. Reduction of carbon dioxide to methanol on n- and p-GaAs and p-InP. Effect of crystal face, electrolyte and current density. Journal of the Electrochemical Society, Vol. 130, No. 8, 1983, pp. 1772–1773.10.1149/1.2120090Search in Google Scholar
[68] Zafrir, M., M. Ulman, Y. Zuckerman, and Halmann. Photoelectrochemical reduction of carbon dioxide to formic acid, formaldehyde and methanol on p-gallium arsenide in an aqueous V(II)-V(III) chloride redox system. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 159, No. 2, 1983, pp. 373–389.10.1016/S0022-0728(83)80635-4Search in Google Scholar
[69] Sears, W. M. and S. R. Morrison. Carbon dioxide reduction on gallium arsenide electrodes. The Journal of Physical Chemistry, Vol. 89, No. 15, 1985, pp. 3295–3298.10.1021/j100261a026Search in Google Scholar
[70] Yoneyama, R., K. Sigumura, and S. Kuwabata. Effects of electrolytes on the photoelectrochemical reduction of carbon dioxide at illuminated p-type cadmium telluride and p-type indium phosphide electrodes in aqueous solutions. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 249, 1988, pp. 143–153.10.1016/0022-0728(88)80355-3Search in Google Scholar
[71] Hirota, K., D. A. Tryk, K. Hashimoto, M. Okawa, and A. Fujishima. Photoelectrochemical reduction of CO2 at high current densities at p‐InP electrodes. Journal of The Electrochemical Society, Vol. 145, No. 5, 1998, pp. L82–L84.10.1149/1.1838494Search in Google Scholar
[72] Aurian-Blajeni, B., M. A. Habib, I. Taniguchi, and J. O. Bockris. The study of adsorbed species during the photoassisted reduction of carbon dioxide at a p-CdTe electrode. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 157, No. 2, 1983, pp. 399–404.10.1016/S0022-0728(83)80367-2Search in Google Scholar
[73] Taniguchi, I., B. Aurian-Blajeni, and J. O. Bockris. Photo-aided reduction of carbon dioxide to carbon monoxide. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 157, No. 1, 1983, pp. 179–182.10.1016/S0022-0728(83)80389-1Search in Google Scholar
[74] Taniguchi, I., B. Aurian-Blajeni, and J. O. Bockris. The reduction of carbon dioxide at illuminated p-type semiconductor electrodes in nonaqueous media. Electrochimica Acta, Vol. 29, No. 7, 1984, pp. 923–932.10.1016/0013-4686(84)87137-6Search in Google Scholar
[75] Junfu, L. and C. Baozhu. Photoelectrochemical reduction of carbon dioxide on a p+/p–Si photocathode in aqueous electrolyte. Journal of Electroanalytical Chemistry, Vol. 324, No. 1–2, 1992, pp. 191–200.10.1016/0022-0728(92)80045-6Search in Google Scholar
[76] Ikeda, S., M. Yoshida, and K. Ito. Photoelectrochemical reduction products of carbon dioxide at metal coated p-GaP photocathodes in aqueous electrolytes. Bulletin of The Chemical Society of Japan, Vol. 58, No. 5, 1985, pp. 1353–1357.10.1246/bcsj.58.1353Search in Google Scholar
[77] Hinogami, R., Y. Nakamura, S. Yae, and Y. Nakato. An approach to ideal semiconductor electrodes for efficient photoelectrochemical reduction of carbon dioxide by modification with small metal particles. The Journal of Physical Chemistry B, Vol. 102, No. 6, 1998, pp. 974–980.10.1021/jp972663hSearch in Google Scholar
[78] Kaneco, S., H. Katsumata, T. Suzuki, and K. Ohta. Photoelectrocatalytic reduction of CO2 in LiOH/methanol at metal-modified p-InP electrodes. Applied Catalysis B: Environmental, Vol. 64, No. 1–2, 2006, pp. 139–145.10.1016/j.apcatb.2005.11.012Search in Google Scholar
[79] Hinogami, R., T. Mori, S. Yae, and Y. Nakato. Efficient photoelectrochemical reduction of carbon dioxide on a p-type silicon (p-Si) electrode modified with very small copper particles. Chemistry Letters, Vol. 23, No. 9, 1994, pp. 1725–1728.10.1246/cl.1994.1725Search in Google Scholar
[80] Aurian-Blajeni, B., I. Taniguchi, and J. O. Bockris. Photoelectrochemical reduction of carbon dioxide using polyaniline-coated silicon. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 149, No. 1–2, 1983, pp. 291–293.10.1016/S0022-0728(83)80579-8Search in Google Scholar
[81] Daube, K. A., D. J. Harrison, T. E. Mallouk, A. J. Ricco, S. Chao, M. S. Wrighton, et al. Electrode-confined catalyst systems for use in optical-to-chemical energy conversion. Journal of Photochemistry, Vol. 29, No. 1–2, 1985, pp. 71–88.10.1016/0047-2670(85)87062-3Search in Google Scholar
[82] Taniguchi, Y., H. Yoneyama, and H. Tamura. Photoelectrochemical reduction of carbon dioxide at p-type gallium phosphide electrodes in the presence of crown ether. Bulletin of The Chemical Society of Japan, Vol. 55, No. 7, 1982, pp. 2034–2039.10.1246/bcsj.55.2034Search in Google Scholar
[83] Parkinson, B. A. and P. F. Weaver. Photoelectrochemical pumping of enzymatic CO2 reduction. Nature, Vol. 309, No. 5964, 1984, pp. 148–149.10.1038/309148a0Search in Google Scholar
[84] Beley, M., J. P. Collin, J. P. Sauvage, J. P. Petit, and P. Chartier. Photoassisted electro-reduction of CO2 on p-GaAs in the presence of Ni cyclam2+. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 206, No. 1–2, 1986, pp. 333–339.10.1016/0022-0728(86)90281-0Search in Google Scholar
[85] Bockris, J. O. and J. C. Wass. The photoelectrocatalytic reduction of carbon dioxide. Journal of The Electrochemical Society, Vol. 136, No. 9, 1989, pp. 2521–2528.10.1149/1.2097455Search in Google Scholar
[86] Petit, J. P., P. Chartier, M. Beley, and J. P. Deville. Molecular catalysts in photoelectrochemical cells. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, Vol. 269, No. 2, 1989, pp. 267–281.10.1016/0022-0728(89)85137-XSearch in Google Scholar
[87] Lany, S. Semiconducting transition metal oxides. Journal of Physics: Condensed Matter, Vol. 27, No. 28, 2015, id. 283203.10.1088/0953-8984/27/28/283203Search in Google Scholar PubMed
[88] Monllor-Satoca, D., M. I. Díez-García, T. Lana-Villarreal, and R. Gómez. Photoelectrocatalytic production of solar fuels with semiconductor oxides: materials, activity and modeling. Chemical Communications, Vol. 56, No. 82, 2020, pp. 12272–12289.10.1039/D0CC04387GSearch in Google Scholar
[89] Rao, C. N. R. Transition metal oxides. Annual Review of Physical Chemistry, Vol. 40, No. 1, 1989, pp. 291–326.10.6028/NBS.NSRDS.49Search in Google Scholar
[90] Cox, P. A. Transition metal oxides: An introduction to their electronic structure and properties. Oxford University Press, New York, United States, 2010.Search in Google Scholar
[91] Inoue, T., A. Fujishima, S. Konishi, and K. Honda. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature, Vol. 277, No. 5698, 1979, pp. 637–638.10.1038/277637a0Search in Google Scholar
[92] Rajeshwar, K. Solar energy conversion and environmental remediation using inorganic semiconductor–liquid interfaces: the road traveled and the way forward. The Journal of Physical Chemistry Letters, Vol. 2, No. 11, 2011, pp. 1301–1309.10.1021/jz200396hSearch in Google Scholar PubMed
[93] Hautier, G., A. Miglio, G. Ceder, G. M. Rignanese, and X. Gonze. Identification and design principles of low hole effective mass p-type transparent conducting oxides. Nature Communications, Vol. 4, No. 1, 2013, pp. 1–7.10.1038/ncomms3292Search in Google Scholar PubMed PubMed Central
[94] Luo, J., L. Steier, M. K. Son, M. Schreier, M. T. Mayer, and M. Grätzel. Cu2O nanowire photocathodes for efficient and durable solar water splitting. Nano Letters, Vol. 16, No. 3, 2016, pp. 1848–1857.10.1021/acs.nanolett.5b04929Search in Google Scholar PubMed
[95] Jang, Y. J., J. W. Jang, S. H. Choi, J. Y. Kim, J. H. Kim, D. H. Youn, et al. Tree branch-shaped cupric oxide for highly effective photoelectrochemical water reduction. Nanoscale, Vol. 7, No. 17, 2015, pp. 7624–7631.10.1039/C5NR00208GSearch in Google Scholar
[96] Cardiel, A. C., K. J. McDonald, and K. S. Choi. Electrochemical growth of copper hydroxy double salt films and their conversion to nanostructured p-type CuO photocathodes. Langmuir, Vol. 33, No. 37, 2017, pp. 9262–9270.10.1021/acs.langmuir.7b00588Search in Google Scholar PubMed
[97] Tilley, S. D., M. Schreier, J. Azevedo, M. Stefik, and M. Graetzel. Ruthenium oxide hydrogen evolution catalysis on composite cuprous oxide water-splitting photocathodes. Advanced Functional Materials, Vol. 24, No. 3, 2013, pp. 303–311.10.1002/adfm.201301106Search in Google Scholar
[98] Morales-Guio, C. G., S. D. Tilley, H. Vrubel, M. Grätzel, and X. Hu. Hydrogen evolution from a copper(I) oxide photocathode coated with an amorphous molybdenum sulphide catalyst. Nature Communications, Vol. 5, No. 1, 2014, pp. 1–7.10.1038/ncomms4059Search in Google Scholar PubMed
[99] Won, D. H., C. H. Choi, J. Chung, and S. I. Woo. Photoelectrochemical production of formic acid and methanol from carbon dioxide on metal-decorated CuO/Cu2O-layered thin films under visible light irradiation. Applied Catalysis B: Environmental, Vol. 158–159, 2014, pp. 217–223.10.1016/j.apcatb.2014.04.021Search in Google Scholar
[100] de Brito, J. F., A. R. Araujo, K. Rajeshwar, and M. V. B. Zanoni. Photoelectrochemical reduction of CO2 and Cu/Cu2O films: product distribution and pH effects. Chemical Engineering Journal, Vol. 264, 2015, pp. 302–309.10.1016/j.cej.2014.11.081Search in Google Scholar
[101] Schreier, M., P. Gao, M. T. Mayer, J. Luo, T. Moehl, M. K. Nazeeruddin, et al. Efficient and selective carbon dioxide reduction on low cost protected Cu2O photocathodes using a molecular catalyst. Energy & Environmental Science, Vol. 8, No. 3, 2015, pp. 855–861.10.1039/C4EE03454FSearch in Google Scholar
[102] Schreier, M., J. Luo, P. Gao, T. Moehl, M. T. Mayer, and M. Grätzel. Covalent immobilization of a molecular catalyst on Cu2O photocathodes for CO reduction. Journal of The American Chemical Society, Vol. 138, No. 6, 2016, pp. 1938–1946.10.1021/jacs.5b12157Search in Google Scholar
[103] Landaeta, E., R. A. Masitas, T. B. Clarke, S. Rafacz, D. A. Nelson, M. Isaacs, et al. Copper-oxide-coated silver nanodendrites for photoelectrocatalytic CO2 reduction to acetate at low overpotential. ACS Applied Nano Materials, Vol. 3, No. 4, 2020, pp. 3478–3486.10.1021/acsanm.0c00210Search in Google Scholar
[104] Paracchino, A., V. Laporte, K. Sivula, M. Grätzel, and E. Thimsen. Highly active oxide photocathode for photoelectrochemical water reduction. Nature Materials, Vol. 10, No. 6, 2011, pp. 456–461.10.1038/nmat3017Search in Google Scholar PubMed
[105] Azevedo, J., S. D. Tilley, M. Schreier, M. Stefik, C. Sousa, J. P. Araújo, et al. Tin oxide as stable protective layer for composite cuprous oxide water-splitting photocathodes. Nano Energy, Vol. 24, 2016, pp. 10–16.10.1016/j.nanoen.2016.03.022Search in Google Scholar
[106] Lee, K., S. Lee, H. Cho, S. Jeong, W. D. Kim, S. Lee, et al. Cu+-incorporated TiO2 overlayer on Cu2O nanowire photocathodes for enhanced photoelectrochemical conversion of CO2 to methanol. Journal of Energy Chemistry, Vol. 27, No. 1, 2018, pp. 264–270.10.1016/j.jechem.2017.04.019Search in Google Scholar
[107] Deng, X., R. Li, S. Wu, L. Wang, J. Hu, J. Ma, et al. Metal–organic framework coating enhances the performance of Cu2O in photoelectrochemical CO2 reduction. Journal of The American Chemical Society, Vol. 141, No. 27, 2019, pp. 10924–10929.10.1021/jacs.9b06239Search in Google Scholar PubMed
[108] Li, P., H. Wang, J. Xu, H. Jing, J. Zhang, H. Han, et al. Reduction of CO2 to low carbon alcohols on CuO FCs/Fe2O3 NTs catalyst with photoelectric dual catalytic interfaces. Nanoscale, Vol. 5, No. 23, 2013, pp. 11748–11754.10.1039/c3nr03352jSearch in Google Scholar PubMed
[109] Kecsenovity, E., B. Endrődi, Z. Pápa, K. Hernádi, K. Rajeshwar, and C. Janáky. Decoration of ultra-long carbon nanotubes with Cu2O nanocrystals: a hybrid platform for enhanced photoelectrochemical CO2 reduction. Journal of Materials Chemistry A, Vol. 4, No. 8, 2016, pp. 3139–3147.10.1039/C5TA10457BSearch in Google Scholar
[110] de Brito, J. F. and M. V. B. Zanoni. On the application of Ti/TiO2/CuO n-p junction semiconductor: a case study of electrolyte, temperature and potential influence on CO2 reduction. Chemical Engineering Journal, Vol. 318, 2017, pp. 264–271.10.1016/j.cej.2016.08.033Search in Google Scholar
[111] Septina, W., R. R. Prabhakar, R. Wick, T. Moehl, and S. D. Tilley. Stabilized solar hydrogen production with CuO/CdS heterojunction thin film photocathodes. Chemistry of Materials, Vol. 29, No. 4, 2017, pp. 1735–1743.10.1021/acs.chemmater.6b05248Search in Google Scholar
[112] Li, J. M., C. W. Tsao, M. J. Fang, C. C. Chen, C. W. Liu, and Y. J. Hsu. TiO2-Au-Cu2O photocathodes: Au-mediated Z-scheme charge transfer for efficient solar-driven photoelectrochemical reduction. ACS Applied Nano Materials, Vol. 1, No. 12, 2018, pp. 6843–6853.10.1021/acsanm.8b01678Search in Google Scholar
[113] Jiang, X. X., X. D. Hu, M. Tarek, P. Saravanan, R. Alqadhi, S. Y. Chin, et al. Tailoring the properties of g-C3N4 with CuO for enhanced photoelectrocatalytic CO2 reduction to methanol. Journal of CO2 Utilization, Vol. 40, 2020, id. 101222.10.1016/j.jcou.2020.101222Search in Google Scholar
[114] Tarek, M., K. M. R. Karim, S. Sarmin, H. R. Ong, H. Abdullah, C. K. Cheng, et al. Photoelectrochemical activity of CuO-CdS heterostructured catalyst for CO2 reduction. In IOP Conference Series: Materials Science and Engineering, Vol. 736, No. 4, 2020, id. 042023.10.1088/1757-899X/736/4/042023Search in Google Scholar
[115] Xing, H., E. Lei, Z. Guo, D. Zhao, and Z. Liu. Enhancement in the charge transport and photocorrosion stability of CuO photocathode: the synergistic effect of spatially separated dual-cocatalysts and p-n heterojunction. Chemical Engineering Journal, Vol. 394, 2020, id. 124907.10.1016/j.cej.2020.124907Search in Google Scholar
[116] Bienkowski, K., A. Małolepszy, L. Stobiński, M. Strawski, P. Wróbel, and R. Solarska. Influence of Fermi-level engineering in multi-interface CuO/Cu2O||rGO||h-WO3||rGO|| photoelectrodes on photoelectrochemical CO2 reduction. Energy Technology, Vol. 10, 2022, id. 2100999.10.1002/ente.202100999Search in Google Scholar
[117] Bosman, A. and H. van Daal. Small-polaron versus band conduction in some transition-metal oxides. Advances in Physics, Vol. 19, No. 77, 1970, pp. 1–117.10.1080/00018737000101071Search in Google Scholar
[118] Rettie, A. J. E., W. D. Chemelewski, D. Emin, and C. B. Mullins. Unravelling small-polaron transport in metal oxide photoelectrodes. The Journal of Physical Chemistry Letters, Vol. 7, No. 3, 2016, pp. 471–479.10.1021/acs.jpclett.5b02143Search in Google Scholar PubMed
[119] Marquardt, M. A., N. A. Ashmore, and D. P. Cann. Crystal chemistry and electrical properties of the delafossite structure. Thin Solid Films, Vol. 496, No. 1, 2006, pp. 146–156.10.1016/j.tsf.2005.08.316Search in Google Scholar
[120] Kumar, M., H. Zhao, and C. Persson. Study of band-structure, optical properties and native defects in AIBIIIO2 (AI = Cu or Ag, BIII = Al, Ga or In) delafossites. Semiconductor Science and Technology, Vol. 28, No. 6, 2013, id. 065003.10.1088/0268-1242/28/6/065003Search in Google Scholar
[121] Sheets, W. C., E. Mugnier, A. Barnabé, T. J. Marks, and K. R. Poeppelmeier. Hydrothermal synthesis of delafossite-type oxides. Chemistry of Materials, Vol. 18, No. 1, 2005, pp. 7–20.10.1021/cm051791cSearch in Google Scholar
[122] Read, C. G., Y. Park, and K. S. Choi. Electrochemical synthesis of p-type CuFeO2 electrodes for use in a photoelectrochemical cell. The Journal of Physical Chemistry Letters, Vol. 3, No. 14, 2012, pp. 1872–1876.10.1021/jz300709tSearch in Google Scholar
[123] Gu, J., A. Wuttig, J. W. Krizan, Y. Hu, Z. M. Detweiler, R. J. Cava, et al. Mg-doped CuFeO2 photocathodes for photoelectrochemical reduction of carbon dioxide. The Journal of Physical Chemistry C, Vol. 117, No. 24, 2013, pp. 12415–12422.10.1021/jp402007zSearch in Google Scholar
[124] Prévot, M. S., X. A. Jeanbourquin, W. S. Bourée, F. Abdi, D. Friedrich, R. van de Krol, et al. Evaluating charge carrier transport and surface states in CuFeO2 photocathodes. Chemistry of Materials, Vol. 29, No. 11, 2017, pp. 4952–4962.10.1021/acs.chemmater.7b01284Search in Google Scholar
[125] Omeiri, S., B. Bellal, A. Bouguelia, Y. Bessekhouad, and M. Trari. Electrochemical and photoelectrochemical characterization of CuFeO2 single crystal. Journal of Solid State Electrochemistry, Vol. 13, No. 9, 2008, pp. 1395–1401.10.1007/s10008-008-0703-3Search in Google Scholar
[126] Fugate, E. A., S. Biswas, M. C. Clement, M. Kim, D. Kim, A. Asthagiri, et al. The role of phase impurities and lattice defects on the electron dynamics and photochemistry of CuFeO2 solar photocathodes. Nano Research, Vol. 12, No. 9, 2019, pp. 2390–2399.10.1007/s12274-019-2493-6Search in Google Scholar
[127] Jiang, C. M., S. E. Reyes-Lillo, Y. Liang, Y. S. Liu, G. Liu, F. M. Toma, et al. Electronic structure and performance bottlenecks of CuFeO2 photocathodes. Chemistry of Materials, Vol. 31, No. 7, 2019, pp. 2524–2534.10.1021/acs.chemmater.9b00009Search in Google Scholar
[128] Hiraga, H., T. Makino, T. Fukumura, H. Weng, and M. Kawasaki. Electronic structure of the delafossite-type CuMO2 (M = Sc, Cr, Mn, Fe, and Co): Optical absorption measurements and first-principles calculations. Physical Review B, Vol. 84, No. 4, 2011, id. 041411.10.1103/PhysRevB.84.041411Search in Google Scholar
[129] Husek, J., A. Cirri, S. Biswas, A. Asthagiri, and L. R. Baker. Hole thermalization dynamics facilitate ultrafast spatial charge separation in CuFeO2 solar photocathodes. The Journal of Physical Chemistry C, Vol. 122, No. 21, 2018, pp. 11300–11304.10.1021/acs.jpcc.8b02996Search in Google Scholar
[130] Ito, M., C. Izawa, and T. Watanabe. Direct fabrication of a CuFeO2/Fe photocathode for solar hydrogen production by hydrothermal method. Chemistry Letters, Vol. 46, No. 6, 2017, pp. 814–816.10.1246/cl.170074Search in Google Scholar
[131] Wuttig, A., J. W. Krizan, J. Gu, J. J. Frick, R. J. Cava, and A. B. Bocarsly. The effect of Mg-doping and Cu nonstoichiometry on the photoelectrochemical response of CuFeO2. Journal of Materials Chemistry A, Vol. 5, No. 1, 2017, pp. 165–171.10.1039/C6TA06504JSearch in Google Scholar
[132] Prévot, M. S., N. Guijarro, and K. Sivula. Enhancing the performance of a robust sol-gel-processed p-type delafossite CuFeO2 photocathode for solar water reduction. ChemSusChem, Vol. 8, No. 8, 2015, pp. 1359–1367.10.1002/cssc.201403146Search in Google Scholar
[133] Deng, Z., X. Fang, S. Wu, Y. Zhao, W. Dong, J. Shao, et al. Structure and optoelectronic properties of Mg-doped CuFeO2 thin films prepared by sol–gel method. Journal of Alloys and Compounds, Vol. 577, 2013, pp. 658–662.10.1016/j.jallcom.2013.06.155Search in Google Scholar
[134] Mohamed, H., E. Chikoidze, A. Ratep, A. Elsoud, M. Boshta, and M. Osman. Synthesis of conducting single-phase CuFeO2 thin films by spray pyrolysis technique. Materials Science in Semiconductor Processing, Vol. 107, 2020, id. 104831.10.1016/j.mssp.2019.104831Search in Google Scholar
[135] Barnabé, A., E. Mugnier, L. Presmanes, and P. Tailhades. Preparation of delafossite CuFeO2 thin films by RF-sputtering on conventional glass substrate. Materials Letters, Vol. 60, No. 29–30, 2006, pp. 3468–3470.10.1016/j.matlet.2006.03.033Search in Google Scholar
[136] Hermans, Y., A. Klein, H. P. Sarker, M. N. Huda, H. Junge, T. Toupance, et al. Pinning of the fermi level in CuFeO2 by polaron formation limiting the photovoltage for photochemical water splitting. Advanced Functional Materials, Vol. 30, No. 10, 2020, id. 1910432.10.1002/adfm.201910432Search in Google Scholar
[137] Benko, F. and F. Koffyberg. Opto-electronic properties of p- and n-type delafossite, CuFeO2. Journal of Physics and Chemistry of Solids, Vol. 48, No. 5, 1987, pp. 431–434.10.1016/0022-3697(87)90103-XSearch in Google Scholar
[138] Jiang, T., Y. Zhao, M. Liu, Y. Chen, Z. Xia, and H. Xue. Enhancing the lifetime of photoinduced charge carriers in CuFeO2 nanoplates by hydrothermal doping of Mg for photoelectrochemical water reduction. Physica Status Solidi (a), Vol. 215, No. 14, 2018, id. 1800056.10.1002/pssa.201800056Search in Google Scholar
[139] Jang, Y. J., Y. B. Park, H. E. Kim, Y. H. Choi, S. H. Choi, and J. S. Lee. Oxygen-intercalated CuFeO2 photocathode fabricated by hybrid microwave annealing for efficient solar hydrogen production. Chemistry of Materials, Vol. 28, No. 17, 2016, pp. 6054–6061.10.1021/acs.chemmater.6b00460Search in Google Scholar
[140] Rudradawong, C. and C. Ruttanapun. Effect of excess oxygen for CuFeO2.06 delafossite on thermoelectric and optical properties. Physica B: Condensed Matter, Vol. 526, 2017, pp. 21–27.10.1016/j.physb.2017.09.046Search in Google Scholar
[141] Cheng, X., J. Ding, Y. Wu, H. Liu, and G. Dawson. The photocathodic properties of a Fe2O3 wrapped CuFeO2 layer on ITO glass for water splitting. Chemical Physics, Vol. 513, 2018, pp. 241–245.10.1016/j.chemphys.2018.08.009Search in Google Scholar
[142] Aqaei, F., M. Zare, and A. Shafiekhani. Role of plasmonic Au nanoparticles embedded in the diamond-like carbon overlayer in the performance of CuFeO2 solar photocathodes. Journal of Solid State Electrochemistry, Vol. 25, No. 4, 2021, pp. 1139–1150.10.1007/s10008-020-04876-9Search in Google Scholar
[143] Oh, Y., W. Yang, J. Tan, H. Lee, J. Park, and J. Moon. Boosting visible light harvesting in p‐type ternary oxides for solar‐to‐hydrogen conversion using inverse opal structure. Advanced Functional Materials, Vol. 29, No. 17, 2019, id. 1900194.10.1002/adfm.201900194Search in Google Scholar
[144] Oh, Y., W. Yang, J. Kim, S. Jeong, and J. Moon. Enhanced photocurrent of transparent CuFeO2 photocathodes by self-light-harvesting architecture. ACS Applied Materials & Interfaces, Vol. 9, No. 16, 2017, pp. 14078–14087.10.1021/acsami.7b01208Search in Google Scholar PubMed
[145] Oh, Y., W. Yang, J. Tan, H. Lee, J. Park, and J. Moon. Photoelectrodes based on 2D opals assembled from Cu-delafossite double-shelled microspheres for an enhanced photoelectrochemical response. Nanoscale, Vol. 10, No. 8, 2018, pp. 3720–3729.10.1039/C7NR07351HSearch in Google Scholar PubMed
[146] Prévot, M. S., Y. Li, N. Guijarro, and K. Sivula. Improving charge collection with delafossite photocathodes: a host–guest CuAlO2/CuFeO2 approach. Journal of Materials Chemistry A, Vol. 4, No. 8, 2016, pp. 3018–3026.10.1039/C5TA06336ASearch in Google Scholar
[147] Zhang, L., H. Cao, Y. Lu, H. Zhang, G. Hou, Y. Tang, et al. Effective combination of CuFeO2 with high temperature resistant Nb-doped TiO2 nanotube arrays for CO2 photoelectric reduction. Journal of Colloid and Interface Science, Vol. 568, 2020, pp. 198–206.10.1016/j.jcis.2020.01.082Search in Google Scholar PubMed
[148] Kang, U., S. K. Choi, D. J. Ham, S. M. Ji, W. Choi, D. S. Han, et al. Photosynthesis of formate from CO2 and water at 1% energy efficiency via copper iron oxide catalysis. Energy & Environmental Science, Vol. 8, No. 9, 2015, pp. 2638–2643.10.1039/C5EE01410GSearch in Google Scholar
[149] Yang, X., E. A. Fugate, Y. Mueanngern, and L. R. Baker. Photoelectrochemical CO2 reduction to acetate on iron–copper oxide catalysts. ACS Catalysis, Vol. 7, No. 1, 2016, pp. 177–180.10.1021/acscatal.6b02984Search in Google Scholar
[150] Kang, U. and H. Park. A facile synthesis of CuFeO2 and CuO composite photocatalyst films for the production of liquid formate from CO2 and water over a month. Journal of Materials Chemistry A, Vol. 5, No. 5, 2017, pp. 2123–2131.10.1039/C6TA09378GSearch in Google Scholar
[151] Oh, S., H. Kang, W. Joo, and Y. Joo. Photoelectrochemical CO2 reduction via Cu2O/CuFeO2 hierarchical nanorods photocatalyst. ChemCatChem, Vol. 12, No. 20, 2020, pp. 5185–5191.10.1002/cctc.202001729Search in Google Scholar
[152] Yoon, S. H., U. Kang, H. Park, A. Abdel-Wahab, and D. S. Han. Computational density functional theory study on the selective conversion of CO2 to formate on homogeneously and heterogeneously mixed CuFeO2 and CuO surfaces. Catalysis Today, Vol. 335, 2019, pp. 345–353.10.1016/j.cattod.2018.12.043Search in Google Scholar
[153] Baiano, C., E. Schiavo, C. Gerbaldi, F. Bella, G. Meligrana, G. Talarico, et al. Role of surface defects in CO2 adsorption and activation on CuFeO2 delafossite oxide. Molecular Catalysis, Vol. 496, 2020, id. 111181.10.1016/j.mcat.2020.111181Search in Google Scholar
[154] Chang, X., T. Wang, and J. Gong. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy & Environmental Science, Vol. 9, No. 7, 2016, pp. 2177–2196.10.1039/C6EE00383DSearch in Google Scholar
[155] Bernstein, N. J., S. A. Akhade, and M. K. Janik. Density functional theory study of carbon dioxide electrochemical reduction on the Fe(100) surface. Physical Chemistry Chemical Physics, Vol. 16, No. 27, 2014, pp. 13708–13717.10.1039/C4CP00266KSearch in Google Scholar PubMed
[156] Berglund, S. P., F. F. Abdi, P. Bogdanoff, A. Chemseddine, D. Friedrich, and R. van de Krol. Comprehensive evaluation of CuBi2O4 as a photocathode material for photoelectrochemical water splitting. Chemistry of Materials, Vol. 28, No. 12, 2016, pp. 4231–4242.10.1021/acs.chemmater.6b00830Search in Google Scholar
[157] Cooper, J. K., Z. Zhang, S. Roychoudhury, C. M. Jiang, S. Gul, Y.S. Liu, et al. CuBi2O4: electronic structure, optical properties, and photoelectrochemical performance limitations of the photocathode. Chemistry of Materials, Vol. 33, No. 3, 2021, pp. 934–945.10.1021/acs.chemmater.0c03930Search in Google Scholar
[158] Jung, H. J., Y. Lim, B. U. Choi, H. B. Bae, W. Jung, S. Ryu, et al. Direct identification of antisite cation intermixing and correlation with electronic conduction in CuBi2O4 for photocathodes. ACS Applied Materials & Interfaces, Vol. 12, No. 39, 2020, pp. 43720–43727.10.1021/acsami.0c12491Search in Google Scholar PubMed
[159] Arai, T., Y. Konishi, Y. Iwasaki, H. Sugihara, and K. Sayama. High-throughput screening using porous photoelectrode for the development of visible-light-responsive semiconductors. Journal of Combinatorial Chemistry, Vol. 9, No. 4, 2007, pp. 574–581.10.1021/cc0700142Search in Google Scholar PubMed
[160] Oropeza, F. E., N. Y. Dzade, A. Pons-Martí, Z. Yang, K. H. L. Zhang, N. H. de Leeuw, et al. Electronic structure and interface energetics of CuBi2O4 photoelectrodes. The Journal of Physical Chemistry C, Vol. 124, No. 41, 2020, pp. 22416–22425.10.1021/acs.jpcc.0c08455Search in Google Scholar PubMed PubMed Central
[161] Sharma, G., Z. Zhao, P. Sarker, B. A. Nail, J. Wang, M. N. Huda, et al. Electronic structure, photovoltage, and photocatalytic hydrogen evolution with p-CuBi2O4 nanocrystals. Journal of Materials Chemistry A, Vol. 4, No. 8, 2016, pp. 2936–2942.10.1039/C5TA07040FSearch in Google Scholar
[162] Cao, D., N. Nasori, Z. Wang, Y. Mi, L. Wen, Y. Yang, et al. p-type CuBi2O4: an easily accessible photocathodic material for high-efficiency water splitting. Journal of Materials Chemistry A, Vol. 4, No. 23, 2016, pp. 8995–9001.10.1039/C6TA01234ESearch in Google Scholar
[163] Hossain, M. K., G. F. Samu, K. Gandha, S. Santhanagopalan, J. P. Liu, C. Janáky, et al. Solution combustion synthesis, characterization, and photocatalytic activity of CuBi2O4 and its nanocomposites with CuO and α-Bi2O3. The Journal of Physical Chemistry C, Vol. 121, No. 15, 2017, pp. 8252–8261.10.1021/acs.jpcc.6b13093Search in Google Scholar
[164] Wang, F., W. Septina, A. Chemseddine, F. F. Abdi, D. Friedrich, P. Bogdanoff, et al. Gradient self-doped CuBi2O4 with highly improved charge separation efficiency. Journal of The American Chemical Society, Vol. 139, No. 42, 2017, pp. 15094–15103.10.1021/jacs.7b07847Search in Google Scholar PubMed
[165] Rodríguez-Gutiérrez, I., R. García-Rodríguez, M. Rodríguez-Pérez, A. Vega-Poot, G. Rodríguez Gattorno, B. A. Parkinson, et al. Charge transfer and recombination dynamics at inkjet-printed CuBi2O4 electrodes for photoelectrochemical water splitting. The Journal of Physical Chemistry C, Vol. 122, No. 48, 2018, pp. 27169–27179.10.1021/acs.jpcc.8b07936Search in Google Scholar
[166] Wang, Y., H. Wang, A. R. Woldu, X. Zhang, and T. He. Optimization of charge behavior in nanoporous CuBi2O4 photocathode for photoelectrochemical reduction of CO2. Catalysis Today, Vol. 335, 2019, pp. 388–394.10.1016/j.cattod.2018.12.047Search in Google Scholar
[167] Xu, Y., J. Jian, F. Li, W. Liu, L. Jia, and H Wang. Porous CuBi2O4 photocathodes with rationally engineered morphology and composition towards high-efficiency photoelectrochemical performance. Journal of Materials Chemistry A, Vol. 7, No. 38, 2019, pp. 21997–22004.10.1039/C9TA07892DSearch in Google Scholar
[168] Zhang, Z., S. A. Lindley, R. Dhall, K. Bustillo, W. Han, E. Xie, et al. Beneficial CuO phase segregation in the ternary p-type oxide photocathode CuBi2O4. ACS Applied Energy Materials, Vol. 2, No. 6, 2019, pp. 4111–4117.10.1021/acsaem.9b00297Search in Google Scholar
[169] Hahn, N. T., V. C. Holmberg, B. A. Korgel, and C. B. Mullins. Electrochemical synthesis and characterization of p-CuBi2O4 thin film photocathodes. The Journal of Physical Chemistry C, Vol. 116, No. 10, 2012, pp. 6459–6466.10.1021/jp210130vSearch in Google Scholar
[170] Park, H. S., C. Y. Lee, and E. Reisner. Photoelectrochemical reduction of aqueous protons with a CuO|CuBi2O4 heterojunction under visible light irradiation. Physical Chemistry Chemical Physics, Vol. 16, No. 41, 2014, pp. 22462–22465.10.1039/C4CP03883ESearch in Google Scholar
[171] Li, J., M. Griep, Y. Choi, and D. Chu. Photoelectrochemical overall water splitting with textured CuBi2O4 as a photocathode. Chemical Communications, Vol. 54, No. 27, 2018, pp. 3331–3334.10.1039/C7CC09041BSearch in Google Scholar
[172] Kang, D., J. C. Hill, Y. Park, and K. S. Choi. Photoelectrochemical properties and photostabilities of high surface area CuBi2O4 and Ag-doped CuBi2O4 photocathodes. Chemistry of Materials, Vol. 28, No. 12, 2016, pp. 4331–4340.10.1021/acs.chemmater.6b01294Search in Google Scholar
[173] Wang, F., A. Chemseddine, F. F. Abdi, R. van de Krol, and S. P. Berglund. Spray pyrolysis of CuBi2O4 photocathodes: improved solution chemistry for highly homogeneous thin films. Journal of Materials Chemistry A, Vol. 5, No. 25, 2017, pp. 12838–12847.10.1039/C7TA03009FSearch in Google Scholar
[174] Gottesman, R., I. Levine, M. Schleuning, R. Irani, D. Abou‐Ras, T. Dittrich, et al. Overcoming phase‐purity challenges in complex metal oxide photoelectrodes: a case study of CuBi2O4. Advanced Energy Materials, Vol. 11, No. 11, 2021, id. 2003474.10.1002/aenm.202003474Search in Google Scholar
[175] Zhang, Z., S. A. Lindley, D. Guevarra, K. Kan, A. Shinde, J. M. Gregoire, et al. Fermi level engineering of passivation and electron transport materials for p‐type CuBi2O4 employing a high‐throughput methodology. Advanced Functional Materials, Vol. 30, No. 24, 2020, id. 2000948.10.1002/adfm.202000948Search in Google Scholar
[176] Lamers, M., M. Sahre, M. J. Müller, D. Abou-Ras, R. van de Krol, and F. F. Abdi. Influence of post-deposition annealing on the photoelectrochemical performance of CuBi2O4 thin films. APL Materials, Vol. 8, No. 6, 2020, id. 061101.10.1063/5.0003005Search in Google Scholar
[177] Lee, J., H. Yoon, K. S. Choi, S. Kim, S. Seo, J. Song, et al. Template engineering of CuBi2O4 single‐crystal thin film photocathodes. Small, Vol. 16, No. 39, 2020, id. 2002429.10.1002/smll.202002429Search in Google Scholar
[178] Gottesman, R., A. Song, I. Levine, M. Krause, A. T. M. N. Islam, D. Abou‐Ras, et al. Pure CuBi2O4 photoelectrodes with increased stability by rapid thermal processing of Bi2O3/CuO grown by pulsed laser deposition. Advanced Functional Materials, Vol. 30, No. 21, 2020, id. 1910832.10.1002/adfm.201910832Search in Google Scholar
[179] Zhang, Q., B. Zhai, Z. Lin, X. Zhao, and P. Diao. Dendritic CuBi2O4 array photocathode coated with conformal TiO2 protection layer for efficient and stable photoelectrochemical hydrogen evolution reaction. The Journal of Physical Chemistry C, Vol. 125, No. 3, 2021, pp. 1890–1901.10.1021/acs.jpcc.0c10421Search in Google Scholar
[180] Song, A., P. Plate, A. Chemseddine, F. Wang, F. F. Abdi, M. Wollgarten, et al. Cu:NiO as a hole-selective back contact to improve the photoelectrochemical performance of CuBi2O4 thin film photocathodes. Journal of Materials Chemistry A, Vol. 7, No. 15, 2019, pp. 9183–9194.10.1039/C9TA01489FSearch in Google Scholar
[181] Song, A., I. Levine, R. van de Krol, T. Dittrich, and S. P. Berglund. Revealing the relationship between photoelectrochemical performance and interface hole trapping in CuBi2O4 heterojunction photoelectrodes. Chemical Science, Vol. 11, No. 41, 2020, pp. 11195–11204.10.1039/D0SC03030ASearch in Google Scholar
[182] Oropeza, F. E., B. T. Feleki, K. H. L. Zhang, E. J. M. Hensen, and J. P. Hofmann. Influence of reduced Cu surface states on the photoelectrochemical properties of CuBi2O4. ACS Applied Energy Materials, Vol. 2, No. 9, 2019, pp. 6866–6874.10.1021/acsaem.9b01333Search in Google Scholar
[183] Feng, K., E. M. Akinoglu, F. Bozheyev, L. Guo, M. Jin, X. Wang, et al. Magnetron sputtered copper bismuth oxide photocathodes for solar water reduction. Journal of Physics D: Applied Physics, Vol. 53, No. 49, 2020, id. 495501.10.1088/1361-6463/abaf25Search in Google Scholar
[184] Zhang, Q., B. Zhai, Z. Lin, X. Zhao, and P. Diao. CuO/CuBi2O4 bilayered heterojunction as an efficient photocathode for photoelectrochemical hydrogen evolution reaction. International Journal of Hydrogen Energy, Vol. 46, No. 21, 2021, pp. 11607–11620.10.1016/j.ijhydene.2021.01.050Search in Google Scholar
[185] Monny, S. A., L. Zhang, Z. Wang, B. Luo, M. Konarova, A. Du, et al. Fabricating highly efficient heterostructured CuBi2O4 photocathodes for unbiased water splitting. Journal of Materials Chemistry A, Vol. 8, No. 5, 2020, pp. 2498–2504.10.1039/C9TA10975GSearch in Google Scholar
[186] Malashchonak, I., E. Streltsov, A. Mazanik, O. Korolik, A. Kulak, D. Puzikova, et al. Effective p-type photocurrent sensitization of n-Bi2O3 with p-CuBi2O4 and p-CuO: Z-scheme photoelectrochemical system. Journal of Solid State Electrochemistry, Vol. 24, No. 2, 2020, pp. 401–409.10.1007/s10008-020-04494-5Search in Google Scholar
[187] Yang, F., X. Yu, Z. Liu, J. Niu, T. Zhang, J. Nie, et al. Preparation of Z-scheme CuBi2O4/Bi2O3 nanocomposites using electrospinning and their enhanced photocatalytic performance. Materials Today Communications, Vol. 26, 2021, id. 101735.10.1016/j.mtcomm.2020.101735Search in Google Scholar
[188] Nishikawa, M., S. Yuto, T. Hasegawa, W. Shiroishi, H. Honghao, Y. Nakabayashi, et al. Compositing effects of CuBi2O4 on visible-light responsive photocatalysts. Materials Science in Semiconductor Processing, Vol. 57, 2017, pp. 12–17.10.1016/j.mssp.2016.09.013Search in Google Scholar
[189] Lai, Y. H., K. C. Lin, C. Y. Yen, and B. J. Jiang. A tandem photoelectrochemical water splitting cell consisting of CuBi2O4 and BiVO4 synthesized from a single Bi4O5I2 nanosheet template. Faraday Discussions, Vol. 215, 2019, pp. 297–312.10.1039/C8FD00183ASearch in Google Scholar PubMed
[190] Lee, W. H., J. Kang, H. S. Park, K. M. Nam, and S. K. Cho. Photoelectrochemical response of Au-decorated CuBi2O4 photocathode in bicarbonate solution. Journal of Electroanalytical Chemistry, Vol. 838, 2019, pp. 172–177.10.1016/j.jelechem.2019.02.055Search in Google Scholar
[191] Wang, Y., H. Wang, and T. He. Study on nanoporous CuBi2O4 photocathode coated with TiO2 overlayer for photoelectrochemical CO2 reduction. Chemosphere, Vol. 264, 2021, id. 128508.10.1016/j.chemosphere.2020.128508Search in Google Scholar PubMed
[192] Díaz-García, A. K., T. Lana-Villarreal, and R. Gómez. Sol-gel copper chromium delafossite thin films as stable oxide photocathodes for water splitting. Journal of Materials Chemistry A, Vol. 3, No. 39, 2015, pp. 19683–19687.10.1039/C5TA05227KSearch in Google Scholar
[193] Varga, A., G. F. Samu, and C. Janáky. Rapid synthesis of interconnected CuCrO2 nanostructures: a promising electrode material for photoelectrochemical fuel generation. Electrochimica Acta, Vol. 272, 2018, pp. 22–32.10.1016/j.electacta.2018.03.185Search in Google Scholar
[194] Lekse, J. W., M. K. Underwood, J. P. Lewis, and C. Matranga. Synthesis, characterization, electronic structure, and photocatalytic behavior of CuGaO2 and CuGa1–xFexO2 (x = 0.05, 0.10, 0.15, 0.20) delafossites. The Journal of Physical Chemistry C, Vol. 116, No. 2, 2012, pp. 1865–1872.10.1021/jp2087225Search in Google Scholar
[195] Bredar, A. R. C., M. D. Blanchet, R. B. Comes, and B. H. Farnum. Evidence and influence of copper vacancies in p-type CuGaO2 mesoporous films. ACS Applied Energy Materials, Vol. 2, No. 1, 2018, pp. 19–28.10.1021/acsaem.8b01558Search in Google Scholar
[196] Kumagai, H., G. Sahara, K. Maeda, M. Higashi, R. Abe, and O. Ishitani. Hybrid photocathode consisting of a CuGaO2 p-type semiconductor and a Ru(II)–Re(I) supramolecular photocatalyst: non-biased visible-light-driven CO2 reduction with water oxidation. Chemical Science, Vol. 8, No. 6, 2017, pp. 4242–4249.10.1039/C7SC00940BSearch in Google Scholar
[197] Rezaul Karim, K. M., H. R. Ong, H. Abdullah, A. Yousuf, C. K. Cheng, and M. Rahman Khan. Photoelectrochemical reduction of carbon dioxide to methanol on p-type CuFe2O4 under visible light irradiation. International Journal of Hydrogen Energy, Vol. 43, No. 39, 2018, pp. 18185–18193.10.1016/j.ijhydene.2018.07.174Search in Google Scholar
[198] Tarek, M., K. M. Rezaul Karim, S. M. Sarkar, A. Deb, H. R. Ong, H. Abdullah, et al. Hetero-structure CdS–CuFe2O4 as an efficient visible light active photocatalyst for photoelectrochemical reduction of CO2 to methanol. International Journal of Hydrogen Energy, Vol. 44, No. 48, 2019, pp. 26271–26284.10.1016/j.ijhydene.2019.08.074Search in Google Scholar
[199] Rezaul Karim, K. M., M. Tarek, H. R. Ong, H. Abdullah, A. Yousuf, C. K. Cheng, et al. Photoelectrocatalytic reduction of carbon dioxide to methanol using CuFe2O4 modified with graphene oxide under visible light irradiation. Industrial & Engineering Chemistry Research, Vol. 58, No. 2, 2018, pp. 563–572.10.1021/acs.iecr.8b03569Search in Google Scholar
[200] Karim, K. M., M. Tarek, S. Sarkar, R. Mouras, H. R. Ong, H. Abdullah, et al. Photoelectrocatalytic reduction of CO2 to methanol over CuFe2O4@PANI photocathode. International Journal of Hydrogen Energy. Vol. 46, No. 48, 2021, pp. 24709–24720.10.1016/j.ijhydene.2020.02.195Search in Google Scholar
[201] Thanh Truc, N. T., N. T. Hanh, M. V. Nguyen, N. T. P. le Chi, N. van Noi, D.T. Tran, et al. Novel direct Z-scheme Cu2V2O7/g-C3N4 for visible light photocatalytic conversion of CO2 into valuable fuels. Applied Surface Science, Vol. 457, 2018, pp. 968–974.10.1016/j.apsusc.2018.07.034Search in Google Scholar
[202] Kormányos, A., A. Thomas, M. N. Huda, P. Sarker, J. P. Liu, N. Poudyal, et al. Solution combustion synthesis, characterization, and photoelectrochemistry of CuNb2O6 and ZnNb2O6 nanoparticles. The Journal of Physical Chemistry C, Vol. 120, No. 29, 2016, pp. 16024–16034.10.1021/acs.jpcc.5b12738Search in Google Scholar
[203] Kamimura, S., N. Murakami, T. Tsubota, and T. Ohno. Fabrication and characterization of a p-type Cu3Nb2O8 photocathode toward photoelectrochemical reduction of carbon dioxide. Applied Catalysis B: Environmental, Vol. 174–175, 2015, pp. 471–476.10.1016/j.apcatb.2015.03.034Search in Google Scholar
[204] Park, J. E., Y. Hu, J. W. Krizan, Q. D. Gibson, U. T. Tayvah, A. Selloni, et al. Stable hydrogen evolution from an AgRhO2 photocathode under visible light. Chemistry of Materials, Vol. 30, No. 8, 2018, pp. 2574–2582.10.1021/acs.chemmater.7b04911Search in Google Scholar
[205] Cheng, X., H. Shen, W. Dong, F. Zheng, L. Fang, X. Su, et al. Nano-Au and ferroelectric polarization mediated Si/ITO/BiFeO3 tandem photocathode for efficient H2 production. Advanced Materials Interfaces, Vol. 3, No. 19, 2016, id. 1600485.10.1002/admi.201600485Search in Google Scholar
[206] Das, S., P. Fourmont, D. Benetti, S. G. Cloutier, R. Nechache, Z. M. Wang, et al. High performance BiFeO3 ferroelectric nanostructured photocathodes. The Journal of Chemical Physics, Vol. 153, No. 8, 2020, id. 084705.10.1063/5.0013192Search in Google Scholar PubMed
[207] Ida, S., K. Yamada, T. Matsunaga, H. Hagiwara, Y. Matsumoto, and T. Ishihara. Preparation of p-type CaFe2O4 photocathodes for producing hydrogen from water. Journal of The American Chemical Society, Vol. 132, No. 49, 2010, pp. 17343–17345.10.1021/ja106930fSearch in Google Scholar PubMed
[208] Sekizawa, K., T. Nonaka, T. Arai, and T. Morikawa. Structural improvement of CaFe2O4 by metal doping toward enhanced cathodic photocurrent. ACS Applied materials & interfaces, Vol. 6, No. 14, 2014, pp. 10969–10973.10.1021/am502500ySearch in Google Scholar PubMed
[209] Wheeler, G. P. and K-S. Choi. Investigation of p-type Ca2Fe2O5 as a photocathode for use in a water splitting photoelectrochemical cell. ACS Applied Energy Materials, Vol. 1, No. 9, 2018, pp. 4917–4923.10.1021/acsaem.8b00934Search in Google Scholar
[210] Yoon, J-S., Y-M. Kim, J-W. Lee, and Y-M. Sung. A novel strategy for enhancing photoelectrochemical performance of Ca2Fe2O5 photocathodes: An integrated experimental and DFT-based approach. Applied Surface Science, Vol. 589, 2022, id. 153012.10.1016/j.apsusc.2022.153012Search in Google Scholar
[211] Pawar, G. S. and A. A. Tahir. Unbiased spontaneous solar fuel production using stable LaFeO3 photoelectrode. Scientific Reports, Vol. 8, No. 1, 2018, id. 3501.10.1038/s41598-018-21821-zSearch in Google Scholar PubMed PubMed Central
[212] Pawar, G. S., A. Elikkottil, S. Seetha, K. S. Reddy, B. Pesala, A. A. Tahir, et al. Enhanced photoactivity and hydrogen generation of LaFeO3 photocathode by plasmonic silver nanoparticle incorporation. ACS Applied Energy Materials, Vol. 1, No. 7, 2018, pp. 3449–3456.10.1021/acsaem.8b00628Search in Google Scholar
[213] Wijten, J. H. J., R. P. H. Jong, G. Mul, and B. M. Weckhuysen. Cathodic electro-deposition of Ni-Mo on semiconducting NiFe2O4 for photo-electrochemical hydrogen evolution in alkaline media. ChemSusChem, Vol. 11, No. 8, 2018, pp. 1374–1381.10.1002/cssc.201800112Search in Google Scholar
[214] Wang, L., H. Tang, and Y. Tian. Carbon-shell-decorated p-semiconductor PbMoO4 nanocrystals for efficient and stable photocathode of photoelectrochemical water reduction. Journal of Power Sources, Vol. 319, 2016, pp. 210–218.10.1016/j.jpowsour.2016.04.059Search in Google Scholar
[215] Yuan, J., P. Wang, C. Hao, and G. Yu. Photoelectrochemical reduction of carbon dioxide at CuInS2/graphene hybrid thin film electrode. Electrochimica Acta, Vol. 193, 2016, pp. 1–6.10.1016/j.electacta.2016.02.037Search in Google Scholar
[216] Yuan, J. and Y. Wang. Photoelectrochemical reduction of carbon dioxide to methanol at CuS/CuO/CuInS2 thin film photocathodes. Journal of The Electrochemical Society, Vol. 164, No. 13, 2017, pp. E475–E479.10.1149/2.1301713jesSearch in Google Scholar
[217] An, C., J. Yuan, and J. Zhu. Reduction of CO2 to ethanol on Cu-In/CuInS2 composite thin film photocathode. Journal of The Electrochemical Society, Vol. 165, No. 16, 2018, pp. H1066–H1071.10.1149/2.0891816jesSearch in Google Scholar
[218] Yuan, J., C. Gu, W. Ding, and C. Hao. Photo-electrochemical reduction of carbon dioxide into methanol at CuFeO2 nanoparticle-decorated CuInS2 thin-film photocathodes. Energy & Fuels, Vol. 34, No. 8, 2020, pp. 9914–9922.10.1021/acs.energyfuels.0c02009Search in Google Scholar
[219] de Brito, J. F., M. A. S. Andrade Jr, M. V. B. Zanoni, and L. H. Mascaro. All-solution processed CuGaS2-based photoelectrodes for CO2 reduction. Journal of CO2 Utilization, Vol. 57, 2022, id. 101902.10.1016/j.jcou.2022.101902Search in Google Scholar
[220] Lai, Y., N. B. Watkins, C. Muzzillo, M. Richter, K. Kan, L. Zhou, et al. Molecular coatings improve the selectivity and durability of CO2 reduction chalcogenide photocathodes. ACS Energy Letters, Vol. 7, No. 3, 2022, pp. 1195–1201.10.1021/acsenergylett.1c02762Search in Google Scholar
[221] Hu, Z., J. Gong, Z. Ye, Y. Liu, X. Xiao, and J. C. Yu. Cu(In,Ga)Se2 for selective and efficient photoelectrochemical conversion of CO2 into CO. Journal of Catalysis, Vol. 348, 2020, pp. 88–95.10.1016/j.jcat.2020.02.015Search in Google Scholar
[222] Pati, P. B., Q. Rang, E. Boutin, S. Diring, S. Jobic, N. Barreau, et al. Photocathode functionalized with a molecular cobalt catalyst for selective carbon dioxide reduction in water. Nature Communications, Vol. 11, No. 1, 2020, pp. 1–9.10.1038/s41467-020-17125-4Search in Google Scholar PubMed PubMed Central
[223] Arai, T., S. Tajima, S. Sato, K. Uemura, T. Morikawa, and T. Kajino. Selective CO2 conversion to formate in water using a CZTS photocathode modified with a ruthenium complex polymer. Chemical communications. Vol. 47, No. 47, 2011, pp. 12664–12666.10.1039/c1cc16160aSearch in Google Scholar PubMed
[224] Zhou, S., K. Sun, J. Hung, X. Lu, B. Xie, D. Zhang, et al. Accelerating electron-transfer and tuning product selectivity through surficial vacancy engineering on CZTS/CdS for photoelectrochemical CO2 reduction. Small, Vol. 17, 2021, id. 2100496.10.1002/smll.202100496Search in Google Scholar PubMed
[225] Ikeda, S., S. Fujikawa, T. Harada, T. H. Nguyen, S. Nakanishi, T. Takayama, et al. Photocathode characteristics of a spray-deposited Cu2ZnGeS4 thin film for CO2 reduction in a CO2-saturated aqueous solution. ACS Applied Energy Materials, Vol. 2, No. 9, 2019, pp. 6911–6918.10.1021/acsaem.9b01418Search in Google Scholar
[226] Chen, Y., X. Feng, M. Liu, J. Su, and S. Shen. Towards efficient solar-to-hydrogen conversion: fundamentals and recent progress in copper-based chalcogenide photocathodes. Nanophotonics, Vol. 5, No. 4, 2016, pp. 524–547.10.1515/nanoph-2016-0027Search in Google Scholar
[227] Galante, M. T., P. V. Santiago, V. Y. Yukuhiro, L. A. Silva, N. A. Dos Reis, C. T. Pires, et al. Aminopolysiloxane as Cu2O photocathode overlayer: photocorrosion inhibitor and low overpotential CO2‐to‐formate selectivity promoter. ChemCatChem, Vol. 13, No. 3, 2021, pp. 859–863.10.1002/cctc.202001638Search in Google Scholar
[228] Pan, H. and C. J. Barile. Electrochemical CO2 reduction on polycrystalline copper by modulating proton transfer with fluoropolymer composites. ACS Applied Energy Materials, Vol. 5, No. 4, 2022, pp. S1–S10.10.1021/acsaem.2c00136Search in Google Scholar
[229] Kecsenovity, E., B. Endrődi, P. S. Tóth, Y. Zou, R. AW. Dryfe, K. Rajeshwar, et al. Enhanced photoelectrochemical performance of cuprous oxide/graphene nanohybrids. Journal of The American Chemical Society, Vol. 139, No. 19, 2017, pp. 6682–6692.10.1021/jacs.7b01820Search in Google Scholar PubMed PubMed Central
[230] Raciti, D. and C. Wang. Recent advances in CO2 reduction electrocatalysis on copper. ACS Energy Letters, Vol. 3, No. 7, 2018, pp. 1545–1556.10.1021/acsenergylett.8b00553Search in Google Scholar
[231] Cheng, D., Z-J. Zhao, G. Zhang, P. Yang, L. Li, H. Gao, et al. The nature of active sites for carbon dioxide electroreduction over oxide-derived copper catalysts. Nature Communications, Vol. 12, No. 1, 2021, pp. 1–8.10.1038/s41467-020-20615-0Search in Google Scholar PubMed PubMed Central
[232] Sequeda, I. N., and A. M. Meléndez. Understanding the role of copper vacancies in photoelectrochemical CO2 reduction on cuprous oxide. The Journal of Physical Chemistry Letters, Vol. 13, 2022, pp. 3667–3673.10.1021/acs.jpclett.2c00751Search in Google Scholar PubMed
© 2022 Ian Lorenzo E. Gonzaga and Candy C. Mercado, published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Review Articles
- State of the art, challenges, and emerging trends: Geopolymer composite reinforced by dispersed steel fibers
- A review on the properties of concrete reinforced with recycled steel fiber from waste tires
- Copper ternary oxides as photocathodes for solar-driven CO2 reduction
- Properties of fresh and hardened self-compacting concrete incorporating rice husk ash: A review
- Basic mechanical and fatigue properties of rubber materials and components for railway vehicles: A literature survey
- Research progress on durability of marine concrete under the combined action of Cl− erosion, carbonation, and dry–wet cycles
- Delivery systems in nanocosmeceuticals
- Study on the preparation process and sintering performance of doped nano-silver paste
- Analysis of the interactions between nonoxide reinforcements and Al–Si–Cu–Mg matrices
- Research Articles
- Study on the influence of structural form and parameters on vibration characteristics of typical ship structures
- Deterioration characteristics of recycled aggregate concrete subjected to coupling effect with salt and frost
- Novel approach to improve shale stability using super-amphiphobic nanoscale materials in water-based drilling fluids and its field application
- Research on the low-frequency multiline spectrum vibration control of offshore platforms
- Multiple wide band gaps in a convex-like holey phononic crystal strip
- Response analysis and optimization of the air spring with epistemic uncertainties
- Molecular dynamics of C–S–H production in graphene oxide environment
- Residual stress relief mechanisms of 2219 Al–Cu alloy by thermal stress relief method
- Characteristics and microstructures of the GFRP waste powder/GGBS-based geopolymer paste and concrete
- Development and performance evaluation of a novel environmentally friendly adsorbent for waste water-based drilling fluids
- Determination of shear stresses in the measurement area of a modified wood sample
- Influence of ettringite on the crack self-repairing of cement-based materials in a hydraulic environment
- Multiple load recognition and fatigue assessment on longitudinal stop of railway freight car
- Synthesis and characterization of nano-SiO2@octadecylbisimidazoline quaternary ammonium salt used as acidizing corrosion inhibitor
- Perforated steel for realizing extraordinary ductility under compression: Testing and finite element modeling
- The influence of oiled fiber, freeze-thawing cycle, and sulfate attack on strain hardening cement-based composites
- Perforated steel block of realizing large ductility under compression: Parametric study and stress–strain modeling
- Study on dynamic viscoelastic constitutive model of nonwater reacted polyurethane grouting materials based on DMA
- Mechanical behavior and mechanism investigation on the optimized and novel bio-inspired nonpneumatic composite tires
- Effect of cooling rate on the microstructure and thermal expansion properties of Al–Mn–Fe alloy
- Research on process optimization and rapid prediction method of thermal vibration stress relief for 2219 aluminum alloy rings
- Failure prevention of seafloor composite pipelines using enhanced strain-based design
- Deterioration of concrete under the coupling action of freeze–thaw cycles and salt solution erosion
- Creep rupture behavior of 2.25Cr1Mo0.25V steel and weld for hydrogenation reactors under different stress levels
- Statistical damage constitutive model for the two-component foaming polymer grouting material
- Nano-structural and nano-constraint behavior of mortar containing silica aggregates
- Influence of recycled clay brick aggregate on the mechanical properties of concrete
- Effect of LDH on the dissolution and adsorption behaviors of sulfate in Portland cement early hydration process
- Comparison of properties of colorless and transparent polyimide films using various diamine monomers
- Study in the parameter influence on underwater acoustic radiation characteristics of cylindrical shells
- Experimental study on basic mechanical properties of recycled steel fiber reinforced concrete
- Dynamic characteristic analysis of acoustic black hole in typical raft structure
- A semi-analytical method for dynamic analysis of a rectangular plate with general boundary conditions based on FSDT
- Research on modification of mechanical properties of recycled aggregate concrete by replacing sand with graphite tailings
- Dynamic response of Voronoi structures with gradient perpendicular to the impact direction
- Deposition mechanisms and characteristics of nano-modified multimodal Cr3C2–NiCr coatings sprayed by HVOF
- Effect of excitation type on vibration characteristics of typical ship grillage structure
- Study on the nanoscale mechanical properties of graphene oxide–enhanced shear resisting cement
- Experimental investigation on static compressive toughness of steel fiber rubber concrete
- Study on the stress field concentration at the tip of elliptical cracks
- Corrosion resistance of 6061-T6 aluminium alloy and its feasibility of near-surface reinforcements in concrete structure
- Effect of the synthesis method on the MnCo2O4 towards the photocatalytic production of H2
- Experimental study of the shear strength criterion of rock structural plane based on three-dimensional surface description
- Evaluation of wear and corrosion properties of FSWed aluminum alloy plates of AA2020-T4 with heat treatment under different aging periods
- Thermal–mechanical coupling deformation difference analysis for the flexspline of a harmonic drive
- Frost resistance of fiber-reinforced self-compacting recycled concrete
- High-temperature treated TiO2 modified with 3-aminopropyltriethoxysilane as photoactive nanomaterials
- Effect of nano Al2O3 particles on the mechanical and wear properties of Al/Al2O3 composites manufactured via ARB
- Co3O4 nanoparticles embedded in electrospun carbon nanofibers as free-standing nanocomposite electrodes as highly sensitive enzyme-free glucose biosensors
- Effect of freeze–thaw cycles on deformation properties of deep foundation pit supported by pile-anchor in Harbin
- Temperature-porosity-dependent elastic modulus model for metallic materials
- Effect of diffusion on interfacial properties of polyurethane-modified asphalt–aggregate using molecular dynamic simulation
- Experimental study on comprehensive improvement of shear strength and erosion resistance of yellow mud in Qiang Village
- A novel method for low-cost and rapid preparation of nanoporous phenolic aerogels and its performance regulation mechanism
- In situ bow reduction during sublimation growth of cubic silicon carbide
- Adhesion behaviour of 3D printed polyamide–carbon fibre composite filament
- An experimental investigation and machine learning-based prediction for seismic performance of steel tubular column filled with recycled aggregate concrete
- Effects of rare earth metals on microstructure, mechanical properties, and pitting corrosion of 27% Cr hyper duplex stainless steel
- Application research of acoustic black hole in floating raft vibration isolation system
- Multi-objective parametric optimization on the EDM machining of hybrid SiCp/Grp/aluminum nanocomposites using Non-dominating Sorting Genetic Algorithm (NSGA-II): Fabrication and microstructural characterizations
- Estimating of cutting force and surface roughness in turning of GFRP composites with different orientation angles using artificial neural network
- Displacement recovery and energy dissipation of crimped NiTi SMA fibers during cyclic pullout tests
Articles in the same Issue
- Review Articles
- State of the art, challenges, and emerging trends: Geopolymer composite reinforced by dispersed steel fibers
- A review on the properties of concrete reinforced with recycled steel fiber from waste tires
- Copper ternary oxides as photocathodes for solar-driven CO2 reduction
- Properties of fresh and hardened self-compacting concrete incorporating rice husk ash: A review
- Basic mechanical and fatigue properties of rubber materials and components for railway vehicles: A literature survey
- Research progress on durability of marine concrete under the combined action of Cl− erosion, carbonation, and dry–wet cycles
- Delivery systems in nanocosmeceuticals
- Study on the preparation process and sintering performance of doped nano-silver paste
- Analysis of the interactions between nonoxide reinforcements and Al–Si–Cu–Mg matrices
- Research Articles
- Study on the influence of structural form and parameters on vibration characteristics of typical ship structures
- Deterioration characteristics of recycled aggregate concrete subjected to coupling effect with salt and frost
- Novel approach to improve shale stability using super-amphiphobic nanoscale materials in water-based drilling fluids and its field application
- Research on the low-frequency multiline spectrum vibration control of offshore platforms
- Multiple wide band gaps in a convex-like holey phononic crystal strip
- Response analysis and optimization of the air spring with epistemic uncertainties
- Molecular dynamics of C–S–H production in graphene oxide environment
- Residual stress relief mechanisms of 2219 Al–Cu alloy by thermal stress relief method
- Characteristics and microstructures of the GFRP waste powder/GGBS-based geopolymer paste and concrete
- Development and performance evaluation of a novel environmentally friendly adsorbent for waste water-based drilling fluids
- Determination of shear stresses in the measurement area of a modified wood sample
- Influence of ettringite on the crack self-repairing of cement-based materials in a hydraulic environment
- Multiple load recognition and fatigue assessment on longitudinal stop of railway freight car
- Synthesis and characterization of nano-SiO2@octadecylbisimidazoline quaternary ammonium salt used as acidizing corrosion inhibitor
- Perforated steel for realizing extraordinary ductility under compression: Testing and finite element modeling
- The influence of oiled fiber, freeze-thawing cycle, and sulfate attack on strain hardening cement-based composites
- Perforated steel block of realizing large ductility under compression: Parametric study and stress–strain modeling
- Study on dynamic viscoelastic constitutive model of nonwater reacted polyurethane grouting materials based on DMA
- Mechanical behavior and mechanism investigation on the optimized and novel bio-inspired nonpneumatic composite tires
- Effect of cooling rate on the microstructure and thermal expansion properties of Al–Mn–Fe alloy
- Research on process optimization and rapid prediction method of thermal vibration stress relief for 2219 aluminum alloy rings
- Failure prevention of seafloor composite pipelines using enhanced strain-based design
- Deterioration of concrete under the coupling action of freeze–thaw cycles and salt solution erosion
- Creep rupture behavior of 2.25Cr1Mo0.25V steel and weld for hydrogenation reactors under different stress levels
- Statistical damage constitutive model for the two-component foaming polymer grouting material
- Nano-structural and nano-constraint behavior of mortar containing silica aggregates
- Influence of recycled clay brick aggregate on the mechanical properties of concrete
- Effect of LDH on the dissolution and adsorption behaviors of sulfate in Portland cement early hydration process
- Comparison of properties of colorless and transparent polyimide films using various diamine monomers
- Study in the parameter influence on underwater acoustic radiation characteristics of cylindrical shells
- Experimental study on basic mechanical properties of recycled steel fiber reinforced concrete
- Dynamic characteristic analysis of acoustic black hole in typical raft structure
- A semi-analytical method for dynamic analysis of a rectangular plate with general boundary conditions based on FSDT
- Research on modification of mechanical properties of recycled aggregate concrete by replacing sand with graphite tailings
- Dynamic response of Voronoi structures with gradient perpendicular to the impact direction
- Deposition mechanisms and characteristics of nano-modified multimodal Cr3C2–NiCr coatings sprayed by HVOF
- Effect of excitation type on vibration characteristics of typical ship grillage structure
- Study on the nanoscale mechanical properties of graphene oxide–enhanced shear resisting cement
- Experimental investigation on static compressive toughness of steel fiber rubber concrete
- Study on the stress field concentration at the tip of elliptical cracks
- Corrosion resistance of 6061-T6 aluminium alloy and its feasibility of near-surface reinforcements in concrete structure
- Effect of the synthesis method on the MnCo2O4 towards the photocatalytic production of H2
- Experimental study of the shear strength criterion of rock structural plane based on three-dimensional surface description
- Evaluation of wear and corrosion properties of FSWed aluminum alloy plates of AA2020-T4 with heat treatment under different aging periods
- Thermal–mechanical coupling deformation difference analysis for the flexspline of a harmonic drive
- Frost resistance of fiber-reinforced self-compacting recycled concrete
- High-temperature treated TiO2 modified with 3-aminopropyltriethoxysilane as photoactive nanomaterials
- Effect of nano Al2O3 particles on the mechanical and wear properties of Al/Al2O3 composites manufactured via ARB
- Co3O4 nanoparticles embedded in electrospun carbon nanofibers as free-standing nanocomposite electrodes as highly sensitive enzyme-free glucose biosensors
- Effect of freeze–thaw cycles on deformation properties of deep foundation pit supported by pile-anchor in Harbin
- Temperature-porosity-dependent elastic modulus model for metallic materials
- Effect of diffusion on interfacial properties of polyurethane-modified asphalt–aggregate using molecular dynamic simulation
- Experimental study on comprehensive improvement of shear strength and erosion resistance of yellow mud in Qiang Village
- A novel method for low-cost and rapid preparation of nanoporous phenolic aerogels and its performance regulation mechanism
- In situ bow reduction during sublimation growth of cubic silicon carbide
- Adhesion behaviour of 3D printed polyamide–carbon fibre composite filament
- An experimental investigation and machine learning-based prediction for seismic performance of steel tubular column filled with recycled aggregate concrete
- Effects of rare earth metals on microstructure, mechanical properties, and pitting corrosion of 27% Cr hyper duplex stainless steel
- Application research of acoustic black hole in floating raft vibration isolation system
- Multi-objective parametric optimization on the EDM machining of hybrid SiCp/Grp/aluminum nanocomposites using Non-dominating Sorting Genetic Algorithm (NSGA-II): Fabrication and microstructural characterizations
- Estimating of cutting force and surface roughness in turning of GFRP composites with different orientation angles using artificial neural network
- Displacement recovery and energy dissipation of crimped NiTi SMA fibers during cyclic pullout tests